You are on page 1of 17

REVIEW ARTICLE

published: 12 January 2012


doi: 10.3389/fimmu.2011.00098

Chronic inflammation in cancer development


Gabriele Multhoff 1,2 *, Michael Molls 1 and Jürgen Radons 3
1
Department of Radiation Oncology, Klinikum rechts der Isar, Technische Universität München, Munich, Germany
2
Clinical Cooperation Group Innate Immunity in Tumor Biology, Helmholtz Zentrum München, Munich, Germany
3
multimmune GmbH, Munich, Germany

Edited by: Chronic inflammatory mediators exert pleiotropic effects in the development of cancer.
Cristina Bonorino, Pontificia
On the one hand, inflammation favors carcinogenesis, malignant transformation, tumor
Universidade Catolica do Rio Grande
do Sul, Brazil growth, invasion, and metastatic spread; on the other hand inflammation can stimulate
Reviewed by: immune effector mechanisms that might limit tumor growth. The link between cancer and
Lidija Klampfer, Montefiore Medical inflammation depends on intrinsic and extrinsic pathways. Both pathways result in the
Center and Albert Einstein Cancer activation of transcription factors such as NF-κB, STAT-3, and HIF-1 and in accumulation of
Center, USA
tumorigenic factors in tumor and microenvironment. STAT-3 and NF-κB interact at multiple
Ana Paula Souza, Pontificia
Universidade Católica do Rio Grande levels and thereby boost tumor-associated inflammation which can suppress anti-tumor
do Sul, Brazil immune responses. These factors also promote tumor growth, progression, and metasta-
*Correspondence: tic spread. IL-1, IL-6, TNF, and PGHS-2 are key mediators of an inflammatory milieu by
Gabriele Multhoff , Department of modulating the expression of tumor-promoting factors. In this review we concentrate on
Radiation Oncology, Klinikum rechts
the crucial role of pro-inflammatory mediators in inflammation-driven carcinogenesis and
der Isar, Technische Universität
München, Ismaningerstr. 22, D-81675 outline molecular mechanisms of IL-1 signaling in tumors. In addition, we elucidate the dual
Munich, Germany. roles of stress proteins as danger signals in the development of anti-cancer immunity and
e-mail: gabriele.multhoff@lrz. anti-apoptotic functions.
tu-muenchen.de
Keywords: inflammation, carcinogenesis, tumorigenic factors, heat shock proteins, NF-κB, STAT-3, IL-1

INTRODUCTION an activation of pattern recognition receptors (PRR) that inter-


Based on the presence of leukocytes in cancerous lesions, Rudolf act with pathogen-associated molecular patterns (PAMP). These
Virchow, the founder of cellular pathology, speculated about an receptors comprise to members of the Toll-like receptor (TLR)
association between chronic inflammation and development of family, nucleotide-binding oligomerization domain-like (NOD-
cancer already in 1863 (Virchow, 1863). In line with this observa- like) receptors (NLR), C-type lectin receptors (CLR), triggering
tion, epidemiological studies indicate that apart from hereditary receptors on myeloid cells (TREM), and retinoic acid inducible
predisposition, inflammation serves as a potential risk factor for gene-I-like receptors (RLR; Kawai and Akira, 2011). Binding of
the development of cancer. Nowadays it is generally accepted PAMP to these receptors leads to an initiation of the host’s immune
that up to 25% of human malignancies are related to chronic response by activation of inflammatory cells. The engagement of
inflammation and to viral and bacterial infections (Hussain and PRR triggers the induction of intracellular signaling pathways that
Harris, 2007). Table 1 provides an overview on inflammatory and induce the activation of numerous transcription factors such as
pathogenic conditions that are considered to be associated with NF-κB, STAT, and FOXO. These factors regulate the expression
malignant transformation. of several genes involved in the innate and adaptive immunity
Cancer-related chronic inflammation facilitates unlimited (Akira et al., 2006; Karin et al., 2006). Inadequate pathogen eradi-
replicative potential, independence of growth factors, resistance cation, recurring tissue injury, prolonged inflammatory signaling,
to growth inhibition, escape of programmed cell death, enhanced and failure of anti-inflammatory mechanisms can cause chronic
angiogenesis, tumor extravasation, and metastasis (Hanahan and inflammation which as a result supports tumorigenesis.
Weinberg, 2000). Cancer-related inflammation represents the sev-
enth hallmark in the development of cancer (Colotta et al., 2009). IMPACT OF INFLAMMATION IN TUMORIGENESIS
Persistent microbial infections induced by parasites, bacteria, and Numerous studies provide evidence that chronic inflammation
viruses and physical and/or chemical stimuli can cause inflam- increases the risk of cancer, promotes tumor progression, and
mation (Coussens and Werb, 2002). Bacterial infections follow- supports metastatic spread (Mantovani et al., 2008; Aggarwal
ing surgical removal of primary tumors can promote metastatic and Gehlot, 2009). In the initial phase of tumor development,
growth in mice (Pidgeon et al., 1999) and humans (Taketomi inflammatory mediators such as cytokines, reactive oxygen species
et al., 1997). This process is mediated most likely by endotoxins (ROS), and reactive nitrogen species (RNS) derived from tumor-
altering the critical balance between cell growth and angiogenesis infiltrating immune cells induce epigenetic alterations in pre-
(Pidgeon et al., 1999). Moreover, chronic inflammation induced malignant lesions and silence tumor suppressor genes (Griven-
by non-infectious agents can also contribute to carcinogenesis nikov and Karin, 2010). During tumor promotion, immune cells
and act as a driving force in tumor development. Apart from secrete cytokines and chemokines that act as survival and prolifera-
toxins, oncoproteins and growth factors can affect the host via tion factors for malignant cells. The angiogenic switch is critical for

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 1


Multhoff et al. Inflammation and cancer

Table 1 | Inflammation and their related cancers. that result in the formation of pro-inflammatory factors including
chemokines, cytokines, and PGHS-2. These molecules recruit and
Inductor Inflammation Cancer activate various leukocyte populations such as macrophages, mast
cells, eosinophils, and neutrophils into the tumor microenviron-
Gut pathogens Inflammatory Colorectal cancer
ment like stromal and endothelial cells as well as infiltrating cells.
bowel disease
This concerted action of tumor and micromilieu results in a more
Tobacco smoke Bronchitis Bronchial lung cancer
pronounced generation of inflammatory mediators that drives the
Helicobacter pylori Gastritis Gastric cancer
progression of a positive amplification loop which further triggers
Human papilloma virus Cervicitis Cervical cancer
tumor growth and invasiveness.
Hepatitic B/C virus Hepatitis Hepatocellular
Proto-oncogene activation represents a critical component in
carcinoma
the intrinsic pathway of cancer-related inflammation. In this con-
Bacteria, gall bladder stones Cholecystitis Gall bladder cancer
text, mutations in RAS genes play an important role in tumori-
Tobacco, genetics, alcohol Pancreatitis Pancreatic cancer
genesis. Overall, up to 30% of all human tumors harbor muta-
Epstein-Barr virus Mononucleosis Burkitt’s lymphoma
tions in canonical RAS genes (KRAS, HRAS, NRAS). Remarkably,
Ultraviolet light Sunburn Melanoma
these oncogenic mutations predominantly affect the KRAS locus,
Asbestos fibers Asbestosis Mesothelioma
with oncogenic KRAS mutations being detected in 25–30% of
Gram-uropathogens Schistosomiasis Bladder cancer
all screened tumor samples (Forbes et al., 2011). The high fre-
(Bilharzia)
quency of KRAS mutations and their appearance in early tumor
Gastric acid, alcohol, tobacco Esophagitis Esophageal
stages argue for a causative role of the K-Ras protein in human
adenocarcinoma
tumorigenesis (Fernandez-Medarde and Santos, 2011). More than
30 years ago the founding members of the RAS gene superfamily
an adequate supply of tumor cells with oxygen, nutrition, growth, (HRAS, NRAS, KRAS) were discovered in human tumors as the
and survival factors (Zumsteg and Christofori, 2009). During first proto-oncogenes. Members of the RAS family are crucial for
tumor progression and metastasis, both tumor and immune cells the connection of up-stream signals to down-stream effector path-
produce cytokines and chemokines leading to an increase in ways that are functionally related to cell cycle progression, growth,
cell survival, motility, and invasiveness (DeNardo et al., 2008). migration, cytoskeletal changes, apoptosis, and senescence. In
Epithelial–mesenchymal transition (EMT), a crucial process in tumor cells, activation of mutated RAS is followed by the induc-
tumor invasiveness and metastasis, is also promoted (Yang and tion of several intracellular signaling pathways. Signaling cascades
Weinberg, 2008). EMT refers to the loss of carcinoma epithelial induced by mutated RAS comprise the RAF/MEK/ERK kinase cas-
phenotype and the acquisition of mesenchymal features (Zeis- cade, the PI3K/AKT pathway, and RalGDS proteins (Downward,
berg and Neilson, 2009). The group of Mehta recently found that 2009), the latter belonging to the family of nucleotide-exchange
aberrant tissue transglutaminase (TG2) expression induces EMT factors activating small GTPases such as RalB. Via the exocyst
in epithelial cells (Kumar et al., 2010). This finding, in conjunc- complex, an octameric protein complex implicated in tethering
tion with the observation that inflammatory signals (e.g., TGF-β, of vesicles to membranes (Yamashita et al., 2010), RalB stim-
TNF, and NF-κB) which induce EMT, also induce TG2 expres- ulates the TANK-binding kinase-1 (TBK-1) resulting in NF-κB
sion (Kawata et al., 2011), suggests a possible link between TG2, activation by IκBα phosphorylation. In cancer cells, a constitutive
inflammation, and cancer progression presumably yielding novel activation of this pathway, via chronic RalB activation, restricts the
therapeutic targets for improved patient outcomes. Other typical initiation of apoptosis after oncogenic stress (Chien et al., 2006).
markers of EMT are cadherin-11 and fibroblast-specific protein Beside NF-κB activation, TBK-1 activates the transcription factors
(FSP)-1 which are associated with an increased motility (Zeisberg IRF-3 and IRF-7 (Hacker and Karin, 2006) leading to the pro-
and Neilson, 2009). Twist is necessary to repress the transcription duction of growth and inflammatory mediators. Previously it has
of E-cadherin (Thiery et al., 2009; Zeisberg and Neilson, 2009). been shown that K-Ras is a direct inducer of pro-inflammatory
IL-6 and pro-angiogenic IL-8 required for the initiation of tumor-
PATHWAYS CONNECTING INFLAMMATION AND CANCER associated inflammation and neovascularization and promoting
According to Mantovani et al. (2008), the connection between tumor growth. In these studies knock-down of IL6, genetic abla-
tumorigenesis and inflammation is mediated via intrinsic and tion of the IL6 gene, or treatment with a neutralizing IL-6 antibody
extrinsic pathways. The intrinsic pathway is activated by genetic retarded K-Ras-driven tumorigenesis (Ancrile et al., 2007). Over-
alterations causing inflammation and neoplasia. These alterations expression of oncogenic K-Ras in tumorigenic HeLa cells induced
comprise mutation-driven proto-oncogene activation, chromo- IL-8 secretion, while IL-8 inhibition reduced growth of these cells
somal rearrangement/amplification, and inactivation of tumor and the number of CD31+ cells in a xenograft tumor model (Spar-
suppressor genes. Transformed cells secrete inflammatory media- mann and Bar-Sagi, 2004). Moreover, TBK-1 and NF-κB signaling
tors and thus generate an inflammatory microenvironment. The have been identified as being essential in K-Ras mutant tumors
extrinsic pathway is driven by inflammation or infections that (Barbie et al., 2009). Regarding these observations it was assumed
increase the risk for the development of cancer in organs at that targeting the NF-κB signaling pathway might be effective in
risk such as the prostate, pancreas, colon, lung, and skin. Both treating RAS-mutated tumors (Downward, 2009). Meylan et al.
pathways interfere in tumor cells and induce the activation of (2009) demonstrated that inhibition of the NF-κB pathway in lung
several transcription factors such as NF-κB, STAT-3, and HIF-1 tumors resulted in significantly reduced tumor growth.

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 2


Multhoff et al. Inflammation and cancer

CRITICAL MOLECULES IN CANCER-RELATED INFLAMMATION RANKL in a variety of other cell types, including tumor cells.
Tumor-associated inflammation requires the presence and acti- Breast cancer cells are able to produce RANKL (Park et al., 2003;
vation of inflammatory cells such as macrophages and granulo- Cross et al., 2006) and stimulate osteoclast differentiation when co-
cytes in the tumor microenvironment, formation of inflamma- cultured with bone marrow stromal cells (Park et al., 2003). HIF-
tory mediators by tumor and stromal cells, tumor remodeling, 1α-induced expression of RANKL initiates an increased migration
and angiogenesis (Kundu and Surh, 2008; Colotta et al., 2009). of breast cancer cells via the PI3K/AKT pathway (Tang et al., 2011).
Accumulation of microbial pathogens and tissue necrosis acti- The expression of RANKL in prostate cancer cells was found
vate transcription factors that are necessary for the expression to be associated with an increased appearance of bone metas-
of, e.g., pro-angiogenic factors (IL-8, VEGF), growth factors (IL- tases (Brown et al., 2001; Chen et al., 2006). In head and neck
6, GM-CSF), anti-apoptotic factors (Bcl-XL , c-FLIP), invasion- squamous cell carcinoma RANKL expression promotes EMT and
promoting factors (MMP-2, MMP-7, MMP-9, uPA), inflamma- tumor progression by inducing VEGF-independent angiogenesis
tory enzymes (PGHS-2, LOX), prostaglandins, iNOS, chemokines (Yamada et al., 2011). Moreover, the activity of RANKL was found
(CCL2, CCL20, IL-8), and pro-inflammatory cytokines (IL-1, IL-6, to be involved in the pathophysiology of osteosarcoma (Mori et al.,
IL-23, TNF, TGF-β, EGF) that support the malignant phenotype. 2007a,b), giant cell tumors of the bone (Ng et al., 2010), Paget’s
All molecules mentioned above are regulated by the transcription sarcomas (Sun et al., 2006), and vascular diseases (Hofbauer and
factor NF-κB, a key orchestrator in innate immunity and inflam- Schoppet, 2004). The expression of RANKL increases in response
mation that has emerged as a crucial tumor promoter (Karin, to pro-inflammatory mediators, such as IL-1 (Fernandez et al.,
2006). The presence of constitutively active NF-κB was found to be 2010; Jurado et al., 2010). In fibroblast-like synoviocytes, murine
associated with poor clinical outcome (for an overview see Aggar- osteoblastic, and fibroblastic cells, IL-23 was found to induce an
wal and Gehlot, 2009). NF-κB activation in inflammatory cells up-regulation of RANKL via STAT-3 and NF-κB signaling path-
in response to infectious pathogens, pro-inflammatory mediators ways (Li et al., 2010; Mori et al., 2011). An exposure of these cells to
as well as necrotic cell products results in the generation of sec- pro-inflammatory cytokines such as IL-1, TNF, and IL-6 resulted
retable factors that support growth, survival, and vascularization into a direct or indirect activation of STAT-3 in a feed-forward
of pre-malignant and malignant cells (Karin, 2006). Activation loop. Further evidence for a crucial role of STAT-3 in the regula-
of NF-κB up-regulates cell cycle mediators (cyclin D1, c-Myc), tion of RANKL is shown by Schulze et al. (2010) who found that
anti-apoptotic (c-FLIP, survivin, Bcl-XL ) and adhesion molecules osteolytic prostate cancer cells induce the expression of RANKL
(ICAM-1, ELAM-1, VCAM-17), proteolytic enzymes (e.g., MMP, in a STAT-3/5-dependent manner. Together these data highlight
uPA), and pro-inflammatory factors (PGHS-2, cytokines) that the significance of a STAT-3/5-mediated cytokine production in
promote an invasive phenotype (Aggarwal and Gehlot, 2009). tumor cell migration and the formation of distant metastases.
iNOS is another important inflammatory mediator that causes
the production of NO by macrophages that links chronic inflam- IL-1 AND TNF
mation and tumorigenesis. Elevated levels of NO have been found Elevated levels of IL-1 have been identified in several human
in numerous pre-cancerous and malignant lesions such as Bar- tumor entities such as melanoma, head and neck, colon, lung, and
rett’s mucosa (Wilson et al., 1998), prostate cancer (Aaltoma et al., breast cancer. Overall, patients harboring IL-1-positive tumors
2001), breast cancer (De Paepe et al., 2002), and gastrointestinal have markedly worse prognoses (Lewis et al., 2006). Due to its
carcinomas (Wink et al., 1998; Jaiswal et al., 2001). The iNOS prod- pleiotropic nature, IL-1 promotes tumor growth and metastasis
uct NO contributes to inflammation-associated tumorigenesis by in an autocrine/paracrine manner. IL-1 is produced by tumor,
inducing DNA damage, suppression of DNA repair, modification stromal and endothelial cells, and the host’s infiltrating immune
of oncoproteins, inhibition of apoptosis, promotion of tumor cells (Lewis et al., 2006). Depending on its subcellular location,
growth, angiogenesis, and metastasis as well as suppression of different IL-1 isoforms mediate different functions. Membrane-
anti-tumor immunity (De Paepe et al., 2002). The NO-mediated bound IL-1α which is expressed on malignant cells induces anti-
inhibition of DNA repair enables cells harboring epigenetic alter- tumor immune responses, whereas, intracellular residing precur-
ations to escape from apoptosis. This results in clonal expansion sors of IL-1α control homeostatic functions including gene expres-
of pre-malignant cells and subsequently to carcinogenesis (Sawa sion, differentiation, and cell growth. In contrast, low concen-
and Ohshima, 2006). Furthermore, NO promotes tumor growth trations of secreted IL-1β down-regulate inflammatory responses
by a transactivation of HIF-1α (Sandau et al., 2000), induces the and immune mechanisms, whereas high concentrations promote
expression of pro-angiogenic VEGF (Ravi et al., 2000), and down- inflammation-associated tissue damage and tumor invasiveness
regulates the tumor suppressor protein p53 (Ambs et al., 1998). (Apte et al., 2006). IL-1 can stimulate other cell types to pro-
duce pro-angiogenic and pro-metastatic mediators and thus plays
RANKL an important role in inflammation-associated carcinogenesis (Lin
RANKL, a member of the TNF superfamily of cytokines, was and Karin, 2007; Voronov et al., 2007). In pancreatic cancer IL-1
originally found in T and dendritic cells (DC). RANKL supports confers chemoresistance via an up-regulation of PGHS-2 (Angst
differentiation and survival of effector cells (Anderson et al., 1997). et al., 2008) and promotes angiogenesis during tumor progression
Moreover, it is essential for the differentiation of bone-resorbing (Shchors et al., 2006).
osteoclasts derived from monocyte–macrophage precursors, and IL-1α and IL-1β exert identical agonist actions by binding to the
enables survival and function of mature osteoclasts (Li et al., 2000; IL-1 receptor type I (IL-1RI). After ligation, IL-1/IL-1RI associates
Teitelbaum, 2000). Recent studies documented an expression of with the IL-1 receptor accessory protein (IL-1RAcP) leading to

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 3


Multhoff et al. Inflammation and cancer

activation of intracellular signal transduction cascades. This com- IL-1β is first synthesized as biologically inactive precursor (pro-
plex recruits a number of intracellular adapter molecules including IL-1β) that is further processed by caspase-1, also known as IL-1-
MyD88 (Watters et al., 2007; Gay et al., 2011) to activate signal converting enzyme (ICE), to the mature form, while pro-IL-1α
transduction pathways such as AP-1, p38MAPK, JNK, and NF- is cleaved by calpain. Although IL-1β contributes to growth and
κB (Figure 1). In particular NF-κB provides a mechanistic link metastatic spread in experimental and human cancers, the mol-
between inflammation and tumorigenesis. NF-κB is a major fac- ecular mechanisms regulating the conversion of pro-IL-1β to the
tor which controls apoptosis-based tumor immune surveillance secreted and active cytokine remains to be elucidated.
mechanisms of pre-neoplastic and malignant cells. NF-κB also reg- An elaborate multi-protein complex, the so-called “inflamma-
ulates tumor angiogenesis and invasiveness (Karin, 2006), and may some,” is responsible for the recruitment and activation of caspase-
contribute to chemoresistance of tumor cells (Fahy et al., 2004). 1 (Martinon et al., 2002). Each inflammasome consists of different
A detailed description of the IL-1 signaling pathway is visualized members of the nucleotide oligomerization domain-like receptor
schematically in Figure 1. (NLR) family of proteins. Two of the best characterized human
A third ligand, the naturally occurring IL-1 receptor antagonist inflammasomes are NALP (NACHT, LRR, and pyrin domain-
(IL-1Ra), also binds to IL-1RI and acts as a true receptor antag- containing protein) 1 inflammasome and NALP2/3 inflamma-
onist. Because of its collagenase and prostaglandin-inhibiting some (Franchi et al., 2009). It was shown previously that in sev-
properties, IL-1Ra (anakinra™) is approved for the treatment eral auto-inflammatory diseases constitutive activation of NALP3
of chronic inflammatory diseases including rheumatoid arthritis inflammasome leads to sustained local and systemic inflamma-
(Dinarello, 1996) and systemic onset juvenile idiopathic arthri- tion mediated by IL-1β (Goldbach-Mansky et al., 2006; Lachmann
tis (Hedrich et al., 2011). It has also been identified as being et al., 2009b). Recently, constitutively activated inflammasome was
powerful in reverting IL-1 effects in numerous pathological set- found in human melanoma cells (Okamoto et al., 2010). In this
tings (Dinarello, 1996). Actually, anakinra was successfully used study human melanoma cells from the late stage of the disease
in treating the rare lymphoproliferative disorder Castleman’s dis- spontaneously secrete biologically active IL-1β in the absence of
ease (El-Osta et al., 2010) as well as in myeloma (Lust et al., 2009) exogenous stimuli because of constitutive activation of the inflam-
rendering the use of anakinra and other IL-1-blocking agents such masome and IL-1 receptor (IL-1R) signaling. From these findings
as canakinumab™(anti-IL-1β antibody) or rilonacept™(construct it can be concluded that IL-1-mediated autoinflammation con-
of the two extracellular chains of IL-1RI/IL-1RAcP complex fused tributes to the development and progression of human melanoma
to the Fc segment of IgG) promising therapeutic approaches in suggesting that inhibiting the inflammasome pathway or reducing
human metastatic diseases. The last two agents have been approved IL-1 activity can be a therapeutic option for melanoma patients.
for the treatment of the cryopyrin-associated periodic syndrome The inflammasome also plays a substantial role in environmen-
(CAPS; Hoffman et al., 2008; Lachmann et al., 2009a), a group- tal cancer. It has been shown previously that silica and asbestos
ing of familial cold auto-inflammatory syndrome, Muckle–Wells both activate the NALP3 inflammasome resulting in an increased
syndrome, and neonatal onset multi-inflammatory disease. As IL-1β production and causing lung inflammation (Dostert et al.,
summarized by Dinarello (2010), there are two meaningful reasons 2008). Chronic exposure to asbestos has been identified as being
for the use of IL-1-blocking agents in the treatment of metastatic a high-risk factor for the development of mesothelioma implying
diseases. On the one hand, none of the above mentioned agents a crucial contribution of inflammasome-mediated inflammation
have been found as being associated with any organ toxicities, to the pathogenesis of mesothelioma. Controversely, in an animal
gastrointestinal, or hematological abnormalities. On the other model of colitis-associated cancer (CAC) the NALP3 inflamma-
hand, unlike TNF-blocking agents IL-1-inhibiting treatments lack some was found to be protective against CAC (Allen et al., 2010).
opportunistic infections although routine bacterial and upper air- The NALP3 inflammasome in DC obviously plays a crucial role
way infections are observed. Due to the safety of IL-1 blockage by linking innate and adaptive immune responses against dying
and the availability of the three therapeutics in limiting IL-1 tumors (Ghiringhelli et al., 2009). Based on these observations
actions, clinical trials are encouraged. An NIH trial of anakinra one can hypothesize that constitutively active NALP3 inflamma-
in the treatment of cutaneous melanoma is ongoing because IL- some as can be found in certain tumors produces large amounts
1 plays a pivotal role in angiogenesis by inducing/up-regulating of IL-1 contributing to cancer-related inflammation and thus pro-
pro-angiogenic IL-8 and VEGF contributing to the pathogenesis moting tumor growth and invasiveness, whereas activation of the
of, e.g., multiple melanoma (Dinarello, 2010). inflammasome in tumor-infiltrating immune cells might be ben-
As a pleiotropic cytokine, IL-1 harbors numerous intensify- eficial in inducing anti-tumor immunity. According to Menu and
ing effects on the physiological functions of diverse innate and Vince (2011), the NALP3 inflammasome can be considered as a
immunocompetent cells (Mizel, 1982), IL-12-mediated induction triple-function agent (“the good, the bad, and the ugly”) in human
of Th1 development (Weaver et al., 1988), and induction of Th17 malignancies.
cells (Sutton et al., 2006). IL-1 effects on immune tolerance are also The pleiotropic cytokine TNF plays a dual role in tumorigene-
reported (Nakata et al., 1995). For instance, IL-1β stimulates func- sis. At high concentrations TNF is destructive to tumor vasculature
tion of memory T cells and impairs that of Treg cells (O’Sullivan and induces necrosis. On the other hand, its critical role in chronic
et al., 2006). This brief overview highlights the complexity of the inflammation and its tumor-promoting capacity are well docu-
mechanisms by which IL-1 regulates all types of immune responses mented (Lin and Yeh, 2005; Mocellin et al., 2005). An increased
including tumor cell eradication. expression of TNF was found in human bladder, breast, colorectal,

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 4


Multhoff et al. Inflammation and cancer

FIGURE 1 | IL-1 signaling in the tumor microenvironment. IL-1 is a critical subsequent phosphorylation of IRAK-4 (Cahill and Rogers, 2008). After
molecule in inflammation-associated carcinogenesis produced directly by recruitment of IRAK-1/Tollip to the complex, IRAK-1 is initially phosphorylated
tumor cells or cells of the tumor microenvironment. IL-1 signal transduction is by IRAK-4 (Born et al., 1998; Dunne and O’Neill, 2003). Subsequently, IRAK-1
initiated by binding of either form of IL-1 to IL-1 receptor type I (IL-1RI), which (and possibly IRAK-2) becomes hyperphosphorylated and dissociates into the
undergoes a conformational change allowing the IL-1 receptor accessory cytoplasm where it binds TNF receptor-associated factor 6 (TRAF-6; Cao et al.,
protein (IL-1RAcP) to recognize the ligated IL-1RI. IL-1RAcP does not recognize 1996). IRAK-1 interacts with membrane-bound TAK-binding protein 2 (TAB-2)
IL-1 but represents an essential component in the IL-1 signaling pathway as well as TAK-1/TAB-1 complex (Dower and Qwarnstrom, 2003) followed by
(Wesche et al., 1997b; Radons et al., 2002). The naturally occurring IL-1 translocation of TAB-2 from the plasma membrane to the signalosome and
receptor antagonist (IL-1Ra) also binds to IL-1RI without leading to its subsequent partial activation of TAK-1 by TAB-2. IRAK-1, presumably as dimer
activation. Ligand-mediated heterodimerization of the receptor complex leads or oligomer, enables dimerization of TRAF-6 resulting in its ubiquitination and
to recruitment of dimeric myeloid differentiation protein 88 (MyD88) via its TIR activation. In close proximity to TAB-2, TAK-1 is partially activated followed by
domain (Muzio et al., 1997; Wesche et al., 1997a; Radons et al., 2003) complete activation through polyubiquitinated TRAF-6 (Kishimoto et al., 2000;
followed by complex formation between IRAK-4, MyD88, and IL-1RAcP and (continued)

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 5


Multhoff et al. Inflammation and cancer

FIGURE 1 | Continued cancer development. IL-1 signaling also involves recruitment of PI3-kinase
Martin and Wesche, 2002) enabling activation of numerous signaling (PI3K) to the IL-1 receptor complex via the p85 regulatory subunit of PI3K
cascades. Polyubiquitination of TRAF-6 obviously occurs through IRAK-2 (Reddy et al., 1997) and subsequent activation of AKT/PKB leading to
(Keating et al., 2007). On the one hand, TAK-1 activates certain members of IKK-dependent activation of NF-κB and AP-1 (Cahill and Rogers, 2008).
the MAP kinase family leading to activation of AP-1 and ATF (Ninomiya-Tsuji Receptor ligation can also activate numerous G proteins resulting in
et al., 1999; O’Neill, 2000; Hefler et al., 2005; Blanco et al., 2008) the latter activation of AP-1 and ATF mediated by several MAP kinases and an
augmenting NF-κB-mediated transcription via transactivation (Jefferies and IκBα-independent transactivation of NF-κB (Singh et al., 1999; Jefferies and
O’Neill, 2000; Cahill and Rogers, 2008). On the other hand, TAK-1 O’Neill, 2000). IL-1 signaling finally regulates gene expression of a great
phosphorylates and activates IKK resulting in phosphorylation and variety of tumorigenic factors including pro-angiogenic factors (IL-8, VEGF),
inactivation of IκBα (Wang et al., 2001). Afterward, IκBα dissociates from the growth factors (IL-6, GM-CSF), anti-apoptotic factors (Bcl-XL , c-FLIP),
complex with NF-κB and undergoes proteasomal degradation. After invasion-promoting factors (MMP-2, MMP-7, MMP-9, uPA), inflammatory
phosphorylation, NF-κB translocates to the nucleus and activates enzymes (PGHS-2, LOX), prostaglandins, iNOS, chemokines (CCL2, CCL20,
NF-κB-dependent gene transcription (Chen and Greene, 2004). Inhibitors of IL-8), and pro-inflammatory cytokines (IL-1, IL-6, IL-23, TNF, TGF-β, EGF,
NF-κB activation are indicated that suppress the inflammatory network in RANKL).

and prostate cancer as well as in leukemia and lymphoma (Balk- It turns out that regenerative or anti-inflammatory activities of
will and Mantovani, 2001). TNF is also produced by cells of the IL-6 are mediated by classic signaling whereas pro-inflammatory
tumor microenvironment. Binding of TNF to the TNF receptor responses of interleukin-6 are rather mediated by trans-signaling
1 (TNFR1) activates signaling cascades of NF-κB and c-Jun N- (Rabe et al., 2008). This is important since therapeutic block-
terminal kinase (JNK) which lead to an up-regulation of several ade of IL-6 by the neutralizing anti-IL-6 receptor monoclonal
pro-inflammatory, pro-angiogenic and invasiveness-promoting antibody tocilizumab™ has recently been approved for the treat-
factors, and to the induction of anti-apoptotic molecules such ment of inflammatory diseases. A recently performed clinical trial
as the caspase-8 inhibitor c-FLIP. Activation of NF-κB in response revealed that IL-6 inhibition by tocilizumab retards joint damage
to TNFR1 terminates the activity of JNK (Kamata et al., 2005). progression in patients with rheumatoid arthritis (Smolen et al.,
The ubiquitin ligase Itch is a substrate of JNK that enables the 2011). Interestingly, inhibition of IL-6R-mediated signaling using
degradation of c-FLIP (Chang et al., 2006). Inhibition of JNK via tocilizumab in a xenograft model of oral squamous cell carcinoma
NF-κB-mediated blockage leads to an inactivation of Itch. This (OSCC) suppressed tumor growth and angiogenesis by down-
prevents degradation of c-FLIP and ensures tumor cell survival. regulating VEGF mRNA expression (Shinriki et al., 2009). Clinical
Apart from its role in tumor initiation, TNF promotes angiogene- studies inclusive those mentioned above have shown that inhibi-
sis and impairs immune surveillance by affecting T cell responses tion of IL-6 signaling by tocilizumab is therapeutically effective not
and the activity of macrophages (Elgert et al., 1998). The tumor- only in chronic inflammatory diseases such as rheumatoid arthri-
promoting role of TNF was confirmed in animal models. In the tis (Nishimoto et al., 2004), juvenile idiopathic arthritis (Yokota
absence of TNF mice do not develop hepatocellular carcinoma in et al., 2004), and Crohn’s disease (Ito et al., 2004) but also in
response to cholestatic hepatitis (Pikarsky et al., 2004). TNF and Castleman’s disease (Nishimoto et al., 2005). In all of these dis-
TNFR-deficient mice were also found to be resistant to chemically eases, tocilizumab ameliorates inflammatory manifestations, and
induced carcinogenesis of the skin (Arnott et al., 2004). These normalizes acute phase protein levels. Given its success in treat-
findings indicate that pro-inflammatory activity of TNF functions ing these diseases, tocilizumab may also prove useful in treating
as a pivotal mediator in tumorigenesis (Lin and Karin, 2007). IL-6-related cancers.
Elevated IL-6 levels are found in numerous tumors such as mul-
IL-6 AND PGHS-2 (COX-2) tiple myeloma (Klein et al., 1992), colorectal cancer (Chung and
IL-6 is another NF-κB-regulated pleiotropic pro-inflammatory Chang, 2003), gastric carcinoma (Kai et al., 2005), and Hodgkin
mediator that enables tumor growth and inhibits apoptosis in a lymphoma (Cozen et al., 2004). Moreover, malignant ascites from
variety of human tumors (Rose-John and Schooltink, 2007). In patients with epithelial ovarian cancer was found to contain high
contrast, IL-6 has also been reported as playing a crucial role in levels of IL-6 (Offner et al., 1995). In breast cancer patients high
terminating inflammation (Hudson et al., 2008). IL-6 signaling via IL-6 concentrations induced by an IL6 gene polymorphism cor-
the membrane-bound receptor IL-6Ra is linked to the JAK/STAT relate with poor prognosis (Berger, 2004). A comparison of non-
pathway (predominantly through activation of STAT-3) and leads metastasizing pancreatic cancer, benign prostatic hyperplasia, and
to the expression of genes encoding for anti-apoptotic cell cycle metastasized pancreatic cancer revealed elevated levels of IL-6 in
progression molecules (Lin and Karin, 2007). In contrast, a solu- the latter, more aggressive tumor (Weiss et al., 2011). In this study
ble form of the IL-6R can bind IL-6 with the same affinity as the it was shown that IL-6 leads to an increased expression of uPA and
membrane-bound form and the complex of IL-6 and the soluble VEGF which implies a crucial role of IL-6 in angiogenesis of pan-
IL-6R (sIL6R) can induce signaling in a process called IL-6 trans- creatic tumors. In OSCC lysophosphatidic acid (LPA), a bioactive
signaling (Peters et al., 1998). Because the IL-6R is only sparely lipid with a growth factor-like activity induces the secretion of IL-6
expressed, IL-6 trans-signaling dramatically increases the num- and IL-8 in an NF-κB- and AP-1-dependent manner (Hwang et al.,
ber of potential IL-6 target cells (Rose-John et al., 2006). Animal 2011). Direct stimulation of human osteoblasts with IL-6 and IL-8
models of inflammatory colon cancer suggest that IL-6 trans- induced the expression of RANKL and thereby promotes osteo-
signaling serves as the major pro-inflammatory paradigm of IL-6 clast formation. From these findings it can be concluded that IL-6
signaling under pathophysiologic conditions (Becker et al., 2004). and IL-8 derived from LPA-stimulated OSCC play a crucial role in

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 6


Multhoff et al. Inflammation and cancer

osteogenesis and bone resorption. Inhibition of IL-6 signaling in the induction of STAT-3 activation (Alshamsan, 2011). Therefore,
colon cancer resulted in a reduced tumor growth in mice (Becker tumor-induced p-STAT-3 in DC can be seen as a promising target
et al., 2004; Greten et al., 2004). Incubation of cholangiocarcinoma for colon cancer immunotherapy. In this context, knocking-down
cells with an anti-IL-6-neutralizing antiserum reduced AKT phos- the IL-6 receptor α-chain of DC vaccines significantly enhanced
phorylation and down-regulated the expression of Mcl-1. This the frequency of tumor-specific CD8+ CTL-producing effector
indicates a contribution of IL-6 in the AKT-mediated survival molecules such as IFN-γ, TNF, FasL, perforin, and granzyme B, and
mechanisms (Kobayashi et al., 2005). generated more CD8+ memory T cells, leading to the substantially
PGHS-2, formerly termed as COX-2, has emerged as another prolonged survival of cytotoxic lymphocytes (Tc1) tumor-bearing
pro-inflammatory mediator in tumorigenesis whose expression mice (Hwang et al., 2010).
is mediated by NF-κB. The expression of PGHS-2 is inducible Most of the PGHS-2-induced effects are mediated through its
in response to stimuli such as mitogens, cytokines, growth fac- product PGE2 (Yoshimatsu et al., 2001a,b). Thus, down-regulation
tors, or hormones. PGHS-2 is the rate-limiting enzyme involved of prostaglandins in tumor tissues by PGHS-2 inhibition blocks
in the conversion of arachidonic acid to prostanoids acting as several neoplastic pathways leading to the suppression of tumor
key mediators of inflammation. Aberrant or increased expression growth (Maier et al., 2004). In this context, PGHS-2 inhibitors
of PGHS-2 has been shown to be involved in the pathogenesis of hold promise for cancer chemoprevention. Among them, the
breast, gastric, colorectal, lung, prostate, head/neck, and pancreatic non-steroidal anti-inflammatory drug (NSAID) celecoxib con-
cancer. PGHS-2 affects cell proliferation, apoptosis, angiogenesis, stitutes a potent and specific inhibitor of the inducible human
and metastasis (Lu et al., 2006; Aggarwal and Gehlot, 2009). Over- PGHS-2. Celecoxib interferes with tumor initiation and tumor
expression of PGHS-2 results in the secretion of large amounts of cell growth in vitro and in vivo. Preclinical studies demonstrate
VEGF and therefore, is associated with increased tumor cell inva- promising anti-cancer effects of celecoxib in colorectal, pancreatic
sion and poor prognosis (Raut et al., 2004; Ladetto et al., 2005). as well as head and neck carcinomas. Additionally, celecoxib was
In human basal cell carcinoma cells elevated levels of PGHS-2 found to increase tumor cell sensitivity toward radiochemotherapy
led to an up-regulated expression of the anti-apoptotic mole- (reviewed by Jendrossek, 2011). Celecoxib has also been found to
cules Mcl-1 and Bcl-2, VEGF and basic fibroblast growth factor impair tissue expression of VEGF, tumor angiogenesis, and metas-
(bFGF; Tjiu et al., 2006). In chronic inflammation, endothelial tasis in an experimental model of pancreatic cancer (Wei et al.,
cells express both, acute phase genes and adhesion molecules that 2004). Thus, modulation of PGHS-2 expression may be a promis-
enable recruitment of leukocytes to the site of tissue damage. ing approach in cancer therapy (Jimeno et al., 2006). However,
Moreover, an enhanced production of prostaglandins mediated due to the high toxicity of PGHS-2 inhibitors, the development of
by PGHS-2 augments vasopermeability leading to a more pro- novel components is necessary (Spektor and Fuster, 2005; Lee et al.,
nounced recruitment of leukocytes (Jura et al., 2005). Leukocytes 2007). Randomized clinical trials and meta-analyses reported on
are the main source of RNS and ROS acting as chemical effectors an increased risk for cardiovascular diseases in patients receiving
in inflammation-driven carcinogenesis (Kundu and Surh, 2008). long-term treatment with PGHS-2 inhibitors. These cardiovas-
We identified a constitutively enhanced expression of PGHS-2 in cular adverse effects include myocardial infarction, stroke, and
human pancreatic adenocarcinoma cells that is further increased cardiovascular death/heart failure (summarized by Trelle et al.,
in the presence of IL-1 (Bauer et al., 2009; Hoffmann et al., 2011). 2011). This increased rate of life-threatening cardiovascular side
The constitutive production of PGHS-2 and its key product PGE2 effects led to the withdrawal of the PGHS-2 inhibitors valde-
in the microenvironment of pancreatic carcinomas accounts for coxib and rofecoxib from the market that had been approved by
an enhanced malignancy of pancreatic tumor cells which is caused the United States Food and Drug Association. A patient-pooled
by inhibition of apoptosis, increase in cell proliferation, induction analysis of adjudicated data from 7,950 patients in six placebo-
of angiogenesis, and invasion of malignant cells into surrounding controlled trials demonstrated a dose regimen-related increase in
tissue (Merati et al., 2001; Kong et al., 2002; Garcea et al., 2005). the risk of serious cardiovascular events after a daily administra-
PGHS-2 also induces the expression of MMP-2 (Surh et al., 2001; tion of 400 and 800 mg celecoxib (Solomon et al., 2008). Although
Sansone et al., 2009; Wang et al., 2009b) and pro-inflammatory IL- long-term treatment with high-dose celecoxib can enhance the
6 and IL-1 via PGE2 (Takahashi et al., 2008). PGHS-2-mediated risk for cardiovascular diseases, the drug is still used at lower doses
effects on growth, angiogenesis, invasiveness, and metastasis are in the United States since it is less toxic compared to other PGHS-2
augmented by the IL-1-induced up-regulation of the enzyme by inhibitors (Solomon et al., 2008; Trelle et al., 2011).
forcing the progression of a positive amplification loop triggered A novel approach to overcome the limitations associated with
by PGE2 and IL-6. Of note, it was demonstrated that PGE2 con- the toxicity of PGHS-2 inhibitors is to combine chemical PGHS-
tributes to cancer progression by inhibiting DC differentiation 2 inhibitors at low doses with naturally occurring compounds
and function, acting paradoxically as an immunosuppressive fac- such as the catechin EGCG which is a promising chemopreven-
tor (Muthuswamy et al., 2010; Stock et al., 2011). In cervical cancer, tive agent derived from green tea (summarized by Cerella et al.,
PGE2 was found to induce a cytokine production profile and phe- 2010). Our group investigated the effects of a combinatorial treat-
notypical features of tolerogenic DC suggesting that the altered ment with celecoxib and EGCG on the expression of IL-1-induced
expression of PGE2 might promote carcinogenesis by favoring tumorigenic factors in human pancreatic adenocarcinoma cells.
(pre)cancer immunotolerance (Herfs et al., 2009). In addition, We found that the combined administration of celecoxib and
IL-6 derived from tumor cells or cells of the tumor microenviron- EGCG can induce synergistic cancer preventive effects in pan-
ment was shown to polarize DC toward immune tolerance through creatic cancer cells by down-regulating tumorigenic factors and

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 7


Multhoff et al. Inflammation and cancer

inducing apoptosis (Härdtner et al., 2009). Previous investiga- exerts anti-tumorigenic effects in tumor cells and in the tumor
tions of our group and others revealed anti-proliferative and microenvironment (for a review see Ben-Neriah and Karin,
apoptosis-inducing effects of EGCG and celecoxib in pancreatic 2011). Evidence for a positive association of NF-κB activation
cancer cells (Chen and Zhang, 2007; Inaba et al., 2008; Xu et al., with tumor-associated inflammation came from colitis-associated
2008; Hoffmann et al., 2011). The anti-proliferative properties of colon cancer (Greten et al., 2004; Pikarsky et al., 2004) and
NSAID such as celecoxib are related to effects on the cell cycle hepatitis-associated hepatocellular carcinoma (Pikarsky et al.,
(Xiong, 2004) including changes in gene expression that favor 2004). Colitis-associated colon cancer represents a classical exam-
cell cycle arrest (Yip-Schneider et al., 2001; Tseng et al., 2002). ple for an inflammation-triggered malignancy. NF-κB activation
Interestingly, several studies revealed an anti-proliferative effect in intestinal epithelial cells of this cancer model was found to
of PGHS-2-selective inhibitors not only in PGHS-2-positive but enhance the survival of pre-malignant progenitor cells by inducing
also in PGHS-2-negative pancreatic tumor cells implying that the anti-apoptotic Bcl-XL (Greten et al., 2004; Pikarsky et al., 2004).
inhibitory action of NSAID on cell proliferation can affect both, NF-κB activation in cancer seems to be related, at least in part,
PGHS-2-dependent and -independent pathways (Molina et al., by mutations in components of the signaling cascade or effects of
1999; Yip-Schneider et al., 2001). Numerous investigations also inflammatory factors in the tumor microenvironment that accu-
documented an apoptosis-inducing potential of NSAID (Maier mulate after NF-κB activation (Karin et al., 2002). Transcriptional
et al., 2004; Suganuma et al., 2011). Celecoxib targets several activation of NF-κB leads to the induction of pro-inflammatory
proteins distinct from PGHS-2 that are involved in the control cytokines (e.g., IL-1, IL-6, TNF), chemokines (IL-8), PGHS-2,
of cell survival and cell death including the anti-apoptotic pro- MMP, and adhesion molecules (ICAM-1, VCAM-1). The presence
teins survivin, Mcl-1, and Bcl-2 (Sakoguchi-Okada et al., 2007; of constitutively active NF-κB in most tumors correlate with a poor
Rudner et al., 2010). Further, PGHS-2-independent molecular clinical outcome (Weichert et al., 2007). Moreover, most chemo-
targets of celecoxib comprise the survival kinase AKT/PKB and preventive agents including nutraceuticals derived from different
its up-stream regulator 3-phosphoinositide-dependent kinase-1 sources have the potential to suppress constitutive and inducible
(PDK-1; Belham et al., 1999; Kulp et al., 2004), cyclin-dependent NF-κB activation pathways (Aggarwal and Gehlot, 2009) in order
kinase inhibitors, and cyclins (Grosch et al., 2006), as well as to block chronic inflammation.
the sarcoplasmic/endoplasmic reticulum calcium ATPase SERCA Whilst NF-κB signaling contributes to both, inflammation-
(Johnson et al., 2002). By counteracting these molecules celecoxib driven carcinogenesis and anti-tumor immunity, STAT-3
interferes with the activation status of caspases and finally induces induces cancer-promoting inflammation and restrains anti-tumor
apoptosis. immune responses by counteracting NF-κB-induced expression of
EGCG-mediated effects on apoptosis include caspase-3/-9 acti- anti-tumor Th1 cytokines (IL-12, IFN-γ; Kortylewski et al., 2005;
vation, PARP cleavage, Bax oligomerization, mitochondrial mem- Yu et al., 2007). Furthermore, STAT-3 contributes to the expansion
brane depolarization, direct interaction with anti-apoptotic mem- and development of Treg and Th17 cells (Wang et al., 2009a; Wu
bers of the Bcl-2 family as well as NF-κB inhibition (Lambert et al., 2009). STAT-3 also induces the expression of tumorigenic
et al., 2005; Shimizu et al., 2005; Inaba et al., 2008). Celecoxib mediators (cytokines, pro-angiogenic, and growth factors) and
is reported to interfere, among others, with the NF-κB signaling their corresponding receptors that in turn activate a STAT-3 medi-
pathway (Niederberger et al., 2001; Shirode and Sylvester, 2010) ated immunoregulatory circuit in the tumor microenvironment
providing the basis for the synergism with EGCG. Based on these (Yu et al., 2007). Thus, the constitutive activation of STAT-3, does
findings one can hypothesize that celecoxib in combination with not only promote cancer-related inflammation but also suppresses
EGCG may promote apoptosis directly or indirectly thus altering anti-tumor immune responses (Yu et al., 2009).
the cellular death threshold in tumor cells (Jendrossek, 2011). In As already mentioned, the association of cancer with chronic
previous studies EGCG has been shown to synergistically enhance inflammation is related to intrinsic and extrinsic pathways, both
the effects of TRAIL (Siddiqui et al., 2008) and PGHS-2 inhibitors leading to activation of NF-κB and STAT-3. Similarly to NF-κB,
NS-398 (Adhami et al., 2007) and celecoxib (Basu and Haldar, constitutively active STAT-3 is found in breast, ovarian, prostate,
2009). In a human lung (Suganuma et al., 2011) and prostate can- and brain tumors, leukemia, lymphoma, and multiple myeloma.
cer (Adhami et al., 2007) model the combination of celecoxib and STAT-3 activation results in the modulation of the expression
EGCG increased tumor cell apoptosis and decreased inflamma- of numerous genes that are crucial for maintaining/amplifying
tion. Other natural compounds affecting the PGHS-2 expression tumor-associated inflammation and promoting tumor growth
include the non-flavonoid polyphenols curcumin from turmeric and progression (Yu et al., 2009). The NF-κB family comprises
Curcuma longa, resveratrol from red wine, isoflavone genistein homo- and heterodimeric transcription factors consisting of RelA,
from lupin as well as omega-3 fatty acids from oily fish flaxseeds. c-Rel, RelB, NF-κB1 (p50 and its precursor p105), and NF-
Such a combinatory administration might have future clinical κB2 (p52 and its precursor p100) with RelA-p50 as being the
implications with respect to an adjuvant therapy in cancer patients, most prominent NF-κB transcription factor (Vallabhapurapu and
since it might reduce the adverse effects of high-dose celecoxib as Karin, 2009). Physiologically, NF-κB is sequestered in the cytosol
a monotherapy. by its inhibitory component, IκBα. Upon phosphorylation by the
IκB kinase complex (IKK), IκBα is degraded in an ubiquitin-
INTERACTIONS OF NF-κB AND STAT-3 dependent manner in the proteasome. Then NF-κB translocates
Both, STAT-3 and NF-κB are crucial for cancer-related inflam- into the nucleus where predominantly RelA-p50 up-regulates the
mation. NF-κB does not only mediate tumorigenesis but also expression of Th1 immunostimulatory genes (IL-12, CD40, CD80)

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 8


Multhoff et al. Inflammation and cancer

that are important for the control of microbial infections and These data were confirmed in a xenograft tumor mouse model.
tumor cells (Yu et al., 2007, 2009). STAT-3 opposes the anti-tumor After orthotopic injection of human tumor cells into immuno-
immune responses mediated by NF-κB within a cell. On the one deficient animals the cell surface density of Hsp70 was greater on
hand, STAT-3 is able to inhibit IKK during acute inflammation metastases than on primary tumors (Multhoff et al., 2000; Stangl
and thus attenuates Th1 immune responses (Lee et al., 2009). On et al., 2006). Interestingly, the corresponding normal tissue of the
the other hand, STAT-3 prolongs the nuclear retention of RelA mice was always found to be membrane Hsp70-negative (Stangl
during oncogenic and chronic inflammation by acting as a co- et al., 2011). These findings might be explained by the fact that
transcription factor for RelA thus contributing to the persistent membrane Hsp70 might facilitate metastases, support adherence
activation of NF-κB during chronic inflammation and the malig- of tumor cells to endothelial cells and organs, or might confer
nant process (Yu et al., 2009). It has been shown previously that resistance to an unfavorable milieu during metastasis. In line with
STAT-3 promotes nuclear localization of RelA by acetyltransferase these findings we could show that overall survival of patients with
p300-mediated acetylation affecting the NF-κB/IκBα interaction membrane Hsp70-positive squamous cell carcinomas of the lung
and avoiding its nuclear export (Chen and Greene, 2004). Since and lower rectal carcinomas was significantly reduced compared
STAT-3 is a prerequisite for p300-mediated acetylation of RelA, to those patients with membrane Hsp70-negative tumors (Pfister
constitutive activity of RelA in tumors requires continuous STAT- et al., 2007). Apart from solid tumors also bone marrow samples of
3 signaling (Lee et al., 2009). Accordingly, increased STAT-3 activity patients suffering from acute (AML) and chronic (CML) myeloid
found in tumors preferentially leads to an association of NF-κB leukemia are frequently membrane Hsp70-positive (Gehrmann
with STAT-3 via p300. Due to the NF-κB activating capacity of et al., 2003). Quantitative analysis revealed that ∼15–20% of the
STAT-3 in malignancies constitutive activity of STAT-3 found in total Hsp70 is present in tumor cell membranes (Gehrmann et al.,
tumors preferentially requires RelA (Yu et al., 2009). As stated 2008). The anchorage of Hsp70 within the plasma membrane is
by the authors, this reciprocal relationship is related to the fact most likely mediated by the tumor-specific glycosphingolipid Gb3
that numerous RelA-encoded target gene products function as (Gehrmann et al., 2008). This led us to the hypothesis that mem-
STAT-3 activators (e.g., IL-6, IL-11, IL-17, IL-21, IL-23, PGHS-2). brane Hsp70 might provide an ideal tumor-specific molecule for
Remarkably, expression of IL-6, IL-17, IL-23, and PGHS-2 (all of a targeted immunotherapeutic approach.
them activating STAT-3) depends on STAT-3 as co-transcription Even in the absence of immunogenic peptides, Hsp70, or a
factor for NF-κB. As a consequence, STAT-3 and NF-κB interact at peptide derived thereof in combination with pro-inflammatory
multiple levels and thereby boost tumor-associated inflammation. cytokines such as IL-2 and IL-15 has the capacity to stimulate the
cytolytic activity of NK cells against membrane Hsp70-positive
HEAT SHOCK PROTEINS AND TUMORIGENESIS tumor cells (Multhoff et al., 1997, 2001). The mechanism of tumor
Heat shock proteins (HSP) are highly conserved proteins expressed cell killing has been identified as perforin-independent granzyme
in a wide range of species where they inhabit nearly all cel- B-mediated apoptosis (Gross et al., 2003b). Granzyme B derived
lular and subcellular compartments. Environmental stress (e.g., from activated NK cells specifically binds to membrane Hsp70 on
heat, hypoxia, bacterial infections, heavy metals, oxidative stress, tumor cells and following Hsp70-mediated endocytosis, apoptosis
inflammation) as well as physiological processes (differentiation, is induced (Gross et al., 2003a,b). Hsp70 also has been detected
proliferation, maturation) result in an increased HSP synthesis on tumor-derived exosomes of membrane Hsp70-positive tumors
(Lindquist and Craig, 1988; DeNagel and Pierce, 1992). Intra- (Gastpar et al., 2005). These data suggest that NK cells might
cellular residing HSP protect cells against lethal damage induced be attracted to membrane Hsp70-positive tumors in vivo via the
by environmental stress, and support folding and transport of secretion of Hsp70 surface-positive exosomes. Incubation of NK
newly produced polypeptides and aberrant proteins (Hartl, 1996). cells with Hsp70 protein or a 14mer-peptide derived from the C-
Depending on their intra-/extracellular localization HSP mediate terminus of Hsp70 is accompanied by an up-regulation of activat-
different functions. On the one hand, up-regulated intracellu- ing receptors on NK cells such as CD94/NKG2C, NKG2D, NKp30,
lar HSP levels protect tumor cells from lethal damage induced NKp44, and NKp46 (Gross et al., 2003b,c). Hsp70 membrane-
by environmental stress. On the other hand, membrane-bound positive tumors are thus efficiently eliminated by NK cells that
and extracellular residing HSP with molecular weights of 70 and had been pre-stimulated with low dose IL-2 plus Hsp70 peptide
90 kDa were identified as key regulators of the host’s immune (Multhoff et al., 1999). Adoptive transfer of these TKD-stimulated
system. NK cells in tumor-bearing mice revealed identical results in vivo
A variety of HSP were found on the plasma membrane of tumor (Botzler et al., 1998; Multhoff et al., 2000; Moser et al., 2002). It
cell lines as determined by selective cell surface protein profil- is known that IL-2-activated NK cells are able to induce regres-
ing (Shin et al., 2003). These findings were confirmed by a broad sion of established lung and liver tumors (Schwarz et al., 1989;
screening program of human tumor biopsies in our laboratory Yasumura et al., 1994; Vujanovic et al., 1995; Whiteside et al.,
using cmHsp70.1 mAb (Stangl et al., 2011). Phenotypic analyses 1998). Our group identified a specific migratory capacity of NK
revealed that Hsp70, the major stress-inducible member of the cells toward Hsp70-postive tumor cells and supernatants derived
HSP70 group, is found on the plasma membrane in 50–70% of thereof. The same effect could be observed for the Hsp70 peptide
colon, lung, pancreas, mammary, head and neck, lung, and uro- TKD (Gastpar et al., 2004). From these results we speculated that
genital carcinomas (Multhoff et al., 1995a,b; Chen et al., 2002). killing of Hsp70-positive tumors in vivo might be related to an
Metastases exhibit an elevated Hsp70 membrane density com- enhanced migratory and cytolytic capacity of pre-activated NK
pared to primary tumors in humans (unpublished observation). cells.

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 9


Multhoff et al. Inflammation and cancer

The rapid induction of HSP in response to environmental enzymes, prostaglandins, iNOS as well as chemokines. Among
stress is based on a variety of genetic and biochemical processes them, IL-1, TNF, and IL-6 act as crucial mediators of
referred to as the heat shock response (HSR; Shamovsky and inflammation-driven tumorigenesis forming an inflammatory
Nudler, 2008). The link between HSR and cancer development network in cancer as outlined in Figure 2. These mediators activate
has been emerging since more than 20 years. HSR is regulated the key transcription factors in tumor-associated inflammation:
mainly at the transcription level by heat shock factors (HSF). NF-κB, STAT-3, and HIF-1 impacting any stage of tumorigen-
Among them, HSF-1 is considered as being the key transcrip- esis such as initiation, promotion as well as progression, and
tion factor of stress-inducible HSP (Pirkkala et al., 2001; Akerfelt metastasis. It is noteworthy that NF-κB might be the central
et al., 2010). As a consequence, HSP are over-expressed in a wide player in tumorigenesis. NF-κB is activated by a great vari-
spectrum of human malignancies contributing to tumor growth, ety of lifestyle-related factors including infectious agents, irra-
differentiation, invasiveness, and metastasis and being associated diation, environmental stimuli, tobacco, stress, dietary agents,
with poor prognosis in certain cancer types (Ciocca and Calder- obesity, and alcohol accounting for almost 95% of all cancers
wood, 2005). HSP over-expression in tumor cells plays a pivotal (Aggarwal and Gehlot, 2009). Modern anti-tumor therapies thus
role in tumorigenesis by inhibiting apoptosis and senescence. In aim to suppress NF-κB activation. Most of the chemopreventive
breast cancer, transformation-induced activation of HSF-1 results agents have been found as being able to suppress NF-κB acti-
in an up-regulated expression of Hsp27 and Hsp70 which in turn vation like the selective PGHS-2 inhibitor celecoxib. Moreover,
results in protection against apoptosis (Calderwood, 2010). HSF- lifestyle-related agents derived from different sources including
1 also triggers expression of Hsp90, an essential factor in tumor fruits, legumes, vegetables, grains, spices, and exercise are also
growth due to its ability to chaperone a variety of oncogenic sig- able to inhibit NF-κB leading to suppression of the inflamma-
naling proteins including Her-2/neu and c-Src (Kamal et al., 2003; tory network (Aggarwal and Gehlot, 2009). Clinical and pre-
Neckers and Lee, 2003; Calderwood, 2010). Several studies have clinical studies are conducted to suppress the inflammatory net-
shown that Her-2/neu (c-ErbB-2) is amplified and over-expressed work by the use of, e.g., steroids (dexamethasone, prednisolone),
in many tumors such as breast, ovarian, and gastric adenocarci- TNF inhibitors (thalidomide = thalomid™, anti-TNF antibod-
noma (Hynes and Stern, 1994). Since HSP over-expression also ies such as infliximab = remicade™, etanercept = enbrel™, adal-
protects from drug-related apoptosis (Khaleque et al., 2005), these imumab = humira™), IL-1 inhibitors (anakinra™ = IL-1 recep-
mechanisms highlight the role of HSP in tumor progression and tor antagonist), PGHS-2 inhibitors (celecoxib), NF-κB inhibitors
therapy resistance. (curcumin, EGCG, piperazine), and RANKL inhibitors (deno-
Recent studies indicate an involvement of HSP such as sumab™ = fully human monoclonal anti-RANKL antibody). A
Hsp70/Hsp72 and Hsp90 in the recognition of PAMP by bind- promising approach in cancer therapy also might be targeting
ing to TLR-4 within lipid rafts (Triantafilou and Triantafilou, HSP, because up-regulated Hsp90 and Hsp70 in cancer cells have
2004; Wheeler et al., 2009). Since extracellular residing Hsp70
acts as a danger signal for the immune system (Matzinger, 1998),
this stress protein has been added to the list of “alarmins.”
Endogenous alarmins and exogenous PAMP both comprise the
group of danger-associated molecular patterns (DAMP; Bianchi,
2007). Hsp70, added exogenously to cells stimulates the pro-
duction of pro-inflammatory cytokines TNF, IL-1β, and IL-6 by
antigen presenting cells (Asea et al., 2000a,b, 2002). Extracellu-
lar Hsp70 has also been found to induce IL-8 production in
human bronchial epithelial cells (Chase et al., 2007). In vitro co-
culturing of colon tumor cell spheroids with normal cells caused
a significant tumor grade-dependent increase in IL-6 produc-
tion thereby altering Hsp70 expression (Paduch et al., 2009).
From these observations it can be concluded that Hsp70 may
enhance the impact of tumorigenic mediators in the tumor
microenvironment.

CONCLUDING REMARKS FIGURE 2 | Interactions of inflammatory mediators in tumor cells.


Cancer-related inflammation has emerged as one of the hall- Schematical simplified representation of the complex intracellular signaling
marks of cancer (Hanahan and Weinberg, 2011). In the last network in cancer. Among the tumorigenic factors produced by tumor cells
or cells of the tumor microenvironment, IL-1, TNF, and IL-6 act as crucial
two decades, several tumorigenic factors have been identified
mediators of inflammation-driven tumorigenesis. In particular IL-1 and TNF
as being implicated in inflammation-associated carcinogenesis. are major pleiotropic cytokines involved in tumor/host interactions.
These factors are released by tumor cells or cells of the tumor Nevertheless, these cytokines function as autocrine growth factors and
microenvironment such as stromal cells, endothelial cells, or modulate the expression of several tumorigenic factors at any stage of
host infiltrating cells, respectively, and include pro-inflammatory tumorigenesis not only affecting proliferation, migration, and survival of
tumor cells but also angiogenesis, invasiveness, metastasis as well as
cytokines, pro-angiogenic and growth-promoting factors, chemoresistance of tumors.
anti-apoptotic and invasion-promoting factors, inflammatory

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 10


Multhoff et al. Inflammation and cancer

been recognized as important drug targets and are under intensive preferred resulting in an enhanced efficacy thereby being less toxic
studies in recent years. HSP are currently being targeted in the ther- to the patient.
apy of breast cancer and other carcinomas and effective drugs for
Hsp90 (e.g., geldanamycin and its analogs) have been synthesized ACKNOWLEDGMENTS
and evaluated in clinical trials (Calderwood and Gong, 2011; Kim The authors want to thank Anett Lange for excellent edi-
and Kim, 2011). HSP vaccines have been intensively studied in the torial assistance. The work was supported by the Deutsche
preceding two decades, proving to be safe and effective in treating Forschungsgemeinschaft (DFG) SFB-824, DFG-Cluster of Excel-
a number of malignancies (Murshid et al., 2011). However, thera- lence: Munich Advanced Photonics, Bundesministerium für Bil-
peutical approaches that completely block one tumorigenic factor dung und Forschung (BMBF) MOBITUM (01EZ0826) Innovative
should be avoided, since it might also interfere with the physio- Therapies (01GU0823), m4 – Personalized Medicine and Targeted
logical anti-tumor immune response and could therefore prove Therapies (01EX1021C), Kompetenzverbund Strahlenforschung
to be harmful. Instead, partial inhibition of numerous factors is (03NUK007E), and multimmune GmbH.

REFERENCES of NOS2-expressing human carci- and cytokine. Nat. Med. 6, 435–442. Belham, C., Wu, S., and Avruch,
Aaltoma, S. H., Lipponen, P. K., and noma cells. Nat. Med. 4, 1371–1376. Asea, A., Rehli, M., Kabingu, E., Boch, J. (1999). Intracellular signalling:
Kosma, V. M. (2001). Inducible Ancrile, B., Lim, K. H., and Counter, J. A., Bare, O., Auron, P. E., Steven- PDK1 – a kinase at the hub of things.
nitric oxide synthase (iNOS) expres- C. M. (2007). Oncogenic Ras- son, M. A., and Calderwood, S. Curr. Biol. 9, R93–R96.
sion and its prognostic value in induced secretion of IL6 is required K. (2002). Novel signal transduc- Ben-Neriah, Y., and Karin, M. (2011).
prostate cancer. Anticancer Res. 21, for tumorigenesis. Genes Dev. 21, tion pathway utilized by extracellu- Inflammation meets cancer, with
3101–3106. 1714–1719. lar HSP70: role of toll-like receptor NF-kappaB as the matchmaker. Nat.
Adhami, V. M., Malik, A., Zaman, Anderson, D. M., Maraskovsky, E., (TLR) 2 and TLR4. J. Biol. Chem. Immunol. 12, 715–723.
N., Sarfaraz, S., Siddiqui, I. A., Billingsley, W. L., Dougall, W. C., 277, 15028–15034. Berger, F. G. (2004). The interleukin-
Syed, D. N., Afaq, F., Pasha, F. Tometsko, M. E., Roux, E. R., Teepe, Balkwill, F., and Mantovani, A. (2001). 6 gene: a susceptibility factor that
S., Saleem, M., and Mukhtar, H. M. C., DuBose, R. F., Cosman, D., Inflammation and cancer: back to may contribute to racial and eth-
(2007). Combined inhibitory effects and Galibert, L. (1997). A homo- Virchow? Lancet 357, 539–545. nic disparities in breast cancer mor-
of green tea polyphenols and selec- logue of the TNF receptor and its Barbie, D. A., Tamayo, P., Boehm, J. S., tality. Breast Cancer Res. Treat. 88,
tive cyclooxygenase-2 inhibitors on ligand enhance T-cell growth and Kim, S. Y., Moody, S. E., Dunn, I. 281–285.
the growth of human prostate dendritic-cell function. Nature 390, F., Schinzel, A. C., Sandy, P., Meylan, Bianchi, M. E. (2007). DAMPs, PAMPs
cancer cells both in vitro and 175–179. E., Scholl, C., Frohling, S., Chan, E. and alarmins: all we need to know
in vivo. Clin. Cancer Res. 13, Angst, E., Reber, H. A., Hines, O. M., Sos, M. L., Michel, K., Mermel, about danger. J. Leukoc. Biol. 81,
1611–1619. J., and Eibl, G. (2008). Mononu- C., Silver, S. J., Weir, B. A., Reiling, J. 1–5.
Aggarwal, B. B., and Gehlot, P. (2009). clear cell-derived interleukin-1 beta H., Sheng, Q., Gupta, P. B., Wadlow, Blanco, S., Sanz-Garcia, M., Santos, C.
Inflammation and cancer: how confers chemoresistance in pancre- R. C., Le, H., Hoersch, S., Wittner, R., and Lazo, P. A. (2008). Mod-
friendly is the relationship for can- atic cancer cells by upregulation B. S., Ramaswamy, S., Livingston, ulation of interleukin-1 transcrip-
cer patients? Curr. Opin. Pharmacol. of cyclooxygenase-2. Surgery 144, D. M., Sabatini, D. M., Meyerson, tional response by the interaction
9, 351–369. 57–65. M., Thomas, R. K., Lander, E. S., between VRK2 and the JIP1 scaf-
Akerfelt, M., Morimoto, R. I., and Sisto- Apte, R. N., Krelin, Y., Song, X., Mesirov, J. P., Root, D. E., Gilliland, fold protein. PLoS ONE 3, e1660.
nen, L. (2010). Heat shock factors: Dotan, S., Recih, E., Elkabets, M., D. G., Jacks, T., and Hahn, W. C. doi:10.1371/journal.pone.0001660
integrators of cell stress, develop- Carmi, Y., Dvorkin, T., White, R. (2009). Systematic RNA interference Born, T. L., Thomassen, E., Bird,
ment and lifespan. Nat. Rev. Mol. M., Gayvoronsky, L., Segal, S., and reveals that oncogenic KRAS-driven T. A., and Sims, J. E. (1998).
Cell Biol. 11, 545–555. Voronov, E. (2006). Effects of micro- cancers require TBK1. Nature 462, Cloning of a novel receptor sub-
Akira, S., Uematsu, S., and Takeuchi, environment- and malignant cell- 108–112. unit, AcPL, required for interleukin-
O. (2006). Pathogen recognition derived interleukin-1 in carcino- Basu, A., and Haldar, S. (2009). Com- 18 signaling. J. Biol. Chem. 273,
and innate immunity. Cell 124, genesis, tumour invasiveness and binatorial effect of epigallocatechin- 29445–29450.
783–801. tumour-host interactions. Eur. J. 3-gallate and TRAIL on pancreatic Botzler, C., Schmidt, J., Luz, A., Jen-
Allen, I. C., TeKippe, E. M., Woodford, Cancer 42, 751–759. cancer cell death. Int. J. Oncol. 34, nen, L., Issels, R., and Multhoff, G.
R. M., Uronis, J. M., Holl, E. K., Arnott, C. H., Scott, K. A., Moore, R. J., 281–286. (1998). Differential Hsp70 plasma-
Rogers, A. B., Herfarth, H. H., Jobin, Robinson, S. C., Thompson, R. G., Bauer, L., Venz, S., Junker, H., Brandt, membrane expression on primary
C., and Ting, J. P. (2010). The NLRP3 and Balkwill, F. R. (2004). Expres- R., and Radons, J. (2009). Nicoti- human tumors and metastases in
inflammasome functions as a nega- sion of both TNF-alpha receptor namide phosphoribosyltransferase mice with severe combined immun-
tive regulator of tumorigenesis dur- subtypes is essential for optimal skin and prostaglandin H2 synthase 2 are odeficiency. Int. J. Cancer 77,
ing colitis-associated cancer. J. Exp. tumour development. Oncogene 23, up-regulated in human pancreatic 942–948.
Med. 207, 1045–1056. 1902–1910. adenocarcinoma cells after stimula- Brown, J. M., Corey, E., Lee, Z. D.,
Alshamsan, A. (2011). Induc- Asea, A., Kabingu, E., Stevenson, tion with interleukin-1. Int. J. Oncol. True, L. D., Yun, T. J., Tondravi, M.,
tion of tolerogenic dendritic M. A., and Calderwood, S. K. 35, 97–107. and Vessella, R. L. (2001). Osteo-
cells by IL-6-secreting CT26 (2000a). HSP70 peptidembearing Becker, C., Fantini, M. C., Schramm, protegerin and rank ligand expres-
colon carcinoma. Immunophar- and peptide-negative preparations C., Lehr, H. A., Wirtz, S., Nikolaev, sion in prostate cancer. Urology 57,
macol. Immunotoxicol. doi: act as chaperokines. Cell Stress Chap- A., Burg, J., Strand, S., Kiesslich, R., 611–616.
10.3109/08923973.2011.625034. erones 5, 425–431. Huber, S., Ito, H., Nishimoto, N., Cahill, C. M., and Rogers, J. T. (2008).
[Epub ahead of print]. Asea, A., Kraeft, S. K., Kurt-Jones, E. A., Yoshizaki, K., Kishimoto, T., Galle, Interleukin (IL) 1beta induction
Ambs, S., Merriam, W. G., Ogunfusika, Stevenson, M. A., Chen, L. B., Fin- P. R., Blessing, M., Rose-John, S., of IL-6 is mediated by a novel
M. O., Bennett,W. P., Ishibe, N., Hus- berg, R. W., Koo, G. C., and Calder- and Neurath, M. F. (2004). TGF- phosphatidylinositol 3-kinase-
sain, S. P., Tzeng, E. E., Geller, D. wood, S. K. (2000b). HSP70 stimu- beta suppresses tumor progression dependent AKT/IkappaB kinase
A., Billiar, T. R., and Harris, C. C. lates cytokine production through a in colon cancer by inhibition of alpha pathway targeting activa-
(1998). p53 and vascular endothelial CD14-dependant pathway, demon- IL-6 trans-signaling. Immunity 21, tor protein-1. J. Biol. Chem. 283,
growth factor regulate tumor growth strating its dual role as a chaperone 491–501. 25900–25912.

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 11


Multhoff et al. Inflammation and cancer

Calderwood, S. K. (2010). Heat shock immune signaling to tumor cell sur- Dostert, C., Petrilli, V., Van, B. R., Steele, carcinogenesis and the implications
proteins in breast cancer progression vival. Cell 127, 157–170. C., Mossman, B. T., and Tschopp, for future therapy. Pancreatology 5,
- a suitable case for treatment? Int. J. Chung, Y. C., and Chang, Y. F. (2003). J. (2008). Innate immune activation 514–529.
Hyperthermia 26, 681–685. Serum interleukin-6 levels reflect the through Nalp3 inflammasome sens- Gastpar, R., Gehrmann, M., Bausero, M.
Calderwood, S. K., and Gong, J. (2011). disease status of colorectal cancer. J. ing of asbestos and silica. Science A., Asea, A., Gross, C., Schroeder, J.
Molecular chaperones in mam- Surg. Oncol. 83, 222–226. 320, 674–677. A., and Multhoff, G. (2005). Heat
mary cancer growth and breast Ciocca, D. R., and Calderwood, S. Dower, S. K., and Qwarnstrom, E. shock protein 70 surface-positive
tumor therapy. J. Cell Biochem. doi: K. (2005). Heat shock proteins E. (2003). Signalling networks, tumor exosomes stimulate migra-
10.1002/jcb.23461. [Epub ahead of in cancer: diagnostic, prognostic, inflammation and innate immunity. tory and cytolytic activity of nat-
print]. predictive, and treatment implica- Biochem. Soc. Trans. 31, 1462–1471. ural killer cells. Cancer Res. 65,
Cao, Z., Xiong, J., Takeuchi, M., Kurama, tions. Cell Stress Chaperones 10, Downward, J. (2009). Cancer: a tumour 5238–5247.
T., and Goeddel, D. V. (1996). TRAF6 86–103. gene’s fatal flaws. Nature 462, 44–45. Gastpar, R., Gross, C., Rossbacher,
is a signal transducer for interleukin- Colotta, F., Allavena, P., Sica, A., Gar- Dunne, A., and O’Neill, L. A. (2003). L., Ellwart, J., Riegger, J., and
1. Nature 383, 443–446. landa, C., and Mantovani, A. (2009). The interleukin-1 receptor/Toll-like Multhoff, G. (2004). The cell
Cerella, C., Sobolewski, C., Dicato, M., Cancer-related inflammation, the receptor superfamily: signal trans- surface-localized heat shock protein
and Diederich, M. (2010). Target- seventh hallmark of cancer: links to duction during inflammation and 70 epitope TKD induces migration
ing COX-2 expression by natural genetic instability. Carcinogenesis 30, host defense. Sci. STKE 2003, re3. and cytolytic activity selectively in
compounds: a promising alterna- 1073–1081. Elgert, K. D., Alleva, D. G., and human NK cells. J. Immunol. 172,
tive strategy to synthetic COX-2 Coussens, L. M., and Werb, Z. (2002). Mullins, D. W. (1998). Tumor- 972–980.
inhibitors for cancer chemopreven- Inflammation and cancer. Nature induced immune dysfunction: the Gay, N. J., Gangloff, M., and O’Neill,
tion and therapy. Biochem. Pharma- 420, 860–867. macrophage connection. J. Leukoc. L. A. (2011). What the Myddosome
col. 80, 1801–1815. Cozen, W., Gill, P. S., Ingles, S. A., Biol. 64, 275–290. structure tells us about the initi-
Chang, L., Kamata, H., Solinas, G., Luo, Masood, R., Martinez-Maza, O., El-Osta, H., Janku, F., and Kurzrock, ation of innate immunity. Trends
J. L., Maeda, S., Venuprasad, K., Cockburn, M. G., Gauderman, W. R. (2010). Successful treatment Immunol. 32, 104–109.
Liu, Y. C., and Karin, M. (2006). J., Pike, M. C., Bernstein, L., Nath- of Castleman’s disease with Gehrmann, M., Liebisch, G., Schmitz,
The E3 ubiquitin ligase itch cou- wani, B. N., Salam, M. T., Danley, interleukin-1 receptor antago- G., Anderson, R., Steinem, C.,
ples JNK activation to TNFalpha- K. L., Wang, W., Gage, J., Gundell- nist (Anakinra). Mol. Cancer Ther. De Maio, A., Pockley, G., and
induced cell death by induc- Miller, S., and Mack, T. M. (2004). 9, 1485–1488. Multhoff, G. (2008). Tumor-specific
ing c-FLIP(L) turnover. Cell 124, IL-6 levels and genotype are asso- Fahy, B. N., Schlieman, M. G., Viru- Hsp70 plasma membrane localiza-
601–613. ciated with risk of young adult dachalam, S., and Bold, R. J. tion is enabled by the glycosphin-
Chase, M. A., Wheeler, D. S., Lierl, Hodgkin lymphoma. Blood 103, (2004). Inhibition of AKT abrogates golipid Gb3. PLoS ONE 3, e1925.
K. M., Hughes, V. S., Wong, H. 3216–3221. chemotherapy-induced NF-kappaB doi:10.1371/journal.pone.0001925
R., and Page, K. (2007). Hsp72 Cross, S. S., Harrison, R. F., Balasubra- survival mechanisms: implications Gehrmann, M., Schmetzer, H., Eiss-
induces inflammation and regu- manian, S. P., Lippitt, J. M., Evans, for therapy in pancreatic cancer. J. ner, G., Haferlach, T., Hidde-
lates cytokine production in airway C. A., Reed, M. W., and Holen, I. Am. Coll. Surg. 198, 591–599. mann, W., and Multhoff, G. (2003).
epithelium through a TLR4- and (2006). Expression of receptor acti- Fernandez, M., Pino, A. M., Figueroa, Membrane-bound heat shock pro-
NF-kappaB-dependent mechanism. vator of nuclear factor kappabeta P., and Rodriguez, J. P. (2010). tein 70 (Hsp70) in acute myeloid
J. Immunol. 179, 6318–6324. ligand (RANKL) and tumour necro- The increased expression of recep- leukemia: a tumor specific recogni-
Chen, G., Sircar, K., Aprikian, A., sis factor related, apoptosis induc- tor activator of nuclear-kappaB lig- tion structure for the cytolytic activ-
Potti, A., Goltzman, D., and Rab- ing ligand (TRAIL) in breast can- and (RANKL) of multiple myeloma ity of autologous NK cells. Haema-
bani, S. A. (2006). Expression cer, and their relations with osteo- bone marrow stromal cells is inhib- tologica 88, 474–476.
of RANKL/RANK/OPG in primary protegerin, oestrogen receptor, and ited by the bisphosphonate iban- Ghiringhelli, F., Apetoh, L., Tesniere, A.,
and metastatic human prostate can- clinicopathological variables. J. Clin. dronate. J. Cell. Biochem. 111, Aymeric, L., Ma, Y., Ortiz, C., Ver-
cer as markers of disease stage and Pathol. 59, 716–720. 130–137. maelen, K., Panaretakis, T., Mignot,
functional regulation. Cancer 107, De Paepe, B., Verstraeten, V. M., De Fernandez-Medarde, A., and Santos, E. G., Ullrich, E., Perfettini, J. L.,
289–298. Potter, C. R., and Bullock, G. R. (2011). Ras in cancer and devel- Schlemmer, F., Tasdemir, E., Uhl, M.,
Chen, L., and Zhang, H. Y. (2007). (2002). Increased angiotensin II opmental diseases. Genes Cancer 2, Genin, P., Civas, A., Ryffel, B., Kanel-
Cancer preventive mechanisms of type-2 receptor density in hyper- 344–358. lopoulos, J., Tschopp, J., Andre, F.,
the green tea polyphenol (-)- plasia, DCIS and invasive carci- Forbes, S. A., Bindal, N., Bamford, Lidereau, R., McLaughlin, N. M.,
epigallocatechin-3-gallate. Molecules noma of the breast is paralleled S., Cole, C., Kok, C. Y., Beare, D., Haynes, N. M., Smyth, M. J., Kroe-
12, 946–957. with increased iNOS expression. Jia, M., Shepherd, R., Leung, K., mer, G., and Zitvogel, L. (2009).
Chen, L. F., and Greene, W. C. (2004). Histochem. Cell Biol. 117, 13–19. Menzies, A., Teague, J. W., Camp- Activation of the NLRP3 inflamma-
Shaping the nuclear action of NF- DeNagel, D. C., and Pierce, S. K. (1992). bell, P. J., Stratton, M. R., and some in dendritic cells induces IL-
kappaB. Nat. Rev. Mol. Cell Biol. 5, A case for chaperones in anti- Futreal, P. A. (2011). COSMIC: min- 1beta-dependent adaptive immu-
392–401. gen processing. Immunol. Today 13, ing complete cancer genomes in nity against tumors. Nat. Med. 15,
Chen, X., Tao, Q., Yu, H., Zhang, L., 86–89. the Catalogue of Somatic Mutations 1170–1178.
and Cao, X. (2002). Tumor cell DeNardo, D. G., Johansson, M., and in Cancer. Nucleic Acids Res. 39, Goldbach-Mansky, R., Dailey, N. J.,
membrane-bound heat shock pro- Coussens, L. M. (2008). Immune D945–D950. Canna, S. W., Gelabert, A., Jones,
tein 70 elicits antitumor immunity. cells as mediators of solid tumor Franchi, L., Eigenbrod, T., Munoz- J., Rubin, B. I., Kim, H. J., Brewer,
Immunol. Lett. 84, 81–87. metastasis. Cancer Metastasis Rev. Planillo, R., and Nunez, G. (2009). C., Zalewski, C., Wiggs, E., Hill, S.,
Chien, Y., Kim, S., Bumeister, R., Loo, 27, 11–18. The inflammasome: a caspase-1- Turner, M. L., Karp, B. I., Aksenti-
Y. M., Kwon, S. W., Johnson, C. Dinarello, C. A. (1996). Biologic basis activation platform that regulates jevich, I., Pucino, F., Penzak, S. R.,
L., Balakireva, M. G., Romeo, Y., for interleukin-1 in disease. Blood immune responses and disease Haverkamp, M. H., Stein, L., Adams,
Kopelovich, L., Gale, M. Jr., Yeaman, 87, 2095–2147. pathogenesis. Nat. Immunol. 10, B. S., Moore, T. L., Fuhlbrigge, R. C.,
C., Camonis, J. H., Zhao, Y., and Dinarello, C. A. (2010). Why not treat 241–247. Shaham, B., Jarvis, J. N., O’Neil, K.,
White, M. A. (2006). RalB GTPase- human cancer with interleukin-1 Garcea, G., Dennison, A. R., Stew- Vehe, R. K., Beitz, L. O., Gardner, G.,
mediated activation of the IkappaB blockade? Cancer Metastasis Rev. 29, ard, W. P., and Berry, D. P. (2005). Hannan, W. P., Warren, R. W., Horn,
family kinase TBK1 couples innate 317–329. Role of inflammation in pancreatic W., Cole, J. L., Paul, S. M., Hawkins,

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 12


Multhoff et al. Inflammation and cancer

P. N., Pham, T. H., Snyder, C., Wes- Hedrich, C. M., Bruck, N., Fiebig, B., Hwang, Y. S., Lee, S. K., Park, K. with its anti-tumour effect and car-
ley, R. A., Hoffmann, S. C., Hol- and Gahr, M. (2011). Anakinra: a K., and Chung, W. Y. (2011). diovascular risks. Biochem. J. 366,
land, S. M., Butman, J. A., and Kast- safe and effective first-line treat- Secretion of IL-6 and IL-8 from 831–837.
ner, D. L. (2006). Neonatal-onset ment in systemic onset juvenile idio- lysophosphatidic acid-stimulated Jura, N., Archer, H., and Bar-Sagi, D.
multisystem inflammatory disease pathic arthritis (SoJIA). Rheumatol. oral squamous cell carcinoma (2005). Chronic pancreatitis, pan-
responsive to interleukin-1beta inhi- Int. doi: 10.1007/s00296-011-2249- promotes osteoclastogenesis and creatic adenocarcinoma and the
bition. N. Engl. J. Med. 355, 581–592. 4. [Epub ahead of print]. bone resorption. Oral Oncol. black box in-between. Cell Res. 15,
Greten, F. R., Eckmann, L., Greten, T. Hefler, L. A., Grimm, C., Lantzsch, doi:10.1016/j.oraloncology.2011.08. 72–77.
F., Park, J. M., Li, Z. W., Egan, L. T., Lampe, D., Leodolter, S., Koelbl, 022. [Epub ahead of print]. Jurado, S., Garcia-Giralt, N., Diez-Perez,
J., Kagnoff, M. F., and Karin, M. H., Heinze, G., Reinthaller, A., Hynes, N. E., and Stern, D. F. (1994). The A., Esbrit, P., Yoskovitz, G., Agueda,
(2004). IKKbeta links inflammation Tong-Cacsire, D., Tempfer, C., and biology of erbB-2/neu/HER-2 and L., Urreizti, R., Perez-Edo, L., Salo,
and tumorigenesis in a mouse model Zeillinger, R. (2005). Interleukin- its role in cancer. Biochim. Biophys. G., Mellibovsky, L., Balcells, S., Grin-
of colitis-associated cancer. Cell 118, 1 and interleukin-6 gene polymor- Acta 1198, 165–184. berg, D., and Nogues, X. (2010).
285–296. phisms and the risk of breast cancer Inaba, H., Nagaoka, Y., Kushima, Y., Effect of IL-1beta, PGE(2), and TGF-
Grivennikov, S. I., and Karin, M. in caucasian women. Clin. Cancer Kumagai, A., Matsumoto, Y., Sak- beta1 on the expression of OPG and
(2010). Inflammation and oncoge- Res. 11, 5718–5721. aguchi, M., Baba, K., and Uesato, S. RANKL in normal and osteoporotic
nesis: a vicious connection. Curr. Herfs, M., Herman, L., Hubert, P., Min- (2008). Comparative examination primary human osteoblasts. J. Cell.
Opin. Genet. Dev. 20, 65–71. ner, F., Arafa, M., Roncarati, P., of anti-proliferative activities of Biochem. 110, 304–310.
Grosch, S., Maier, T. J., Schiffmann, Henrotin, Y., Boniver, J., and Del- (-)-epigallocatechin gallate and Kai, H., Kitadai, Y., Kodama, M., Cho,
S., and Geisslinger, G. (2006). venne, P. (2009). High expression of (-)-epigallocatechin against S., Kuroda, T., Ito, M., Tanaka,
Cyclooxygenase-2 (COX-2)- PGE2 enzymatic pathways in cervi- HCT116 colorectal carcinoma S., Ohmoto, Y., and Chayama, K.
independent anticarcinogenic cal (pre)neoplastic lesions and func- cells. Biol. Pharm. Bull. 31, 79–84. (2005). Involvement of proinflam-
effects of selective COX-2 inhibitors. tional consequences for antigen- Ito, H., Takazoe, M., Fukuda, Y., Hibi, matory cytokines IL-1beta and IL-6
J. Natl. Cancer Inst. 98, 736–747. presenting cells. Cancer Immunol. T., Kusugami, K., Andoh, A., Mat- in progression of human gastric car-
Gross, C., Hansch, D., Gastpar, R., and Immunother. 58, 603–614. sumoto, T., Yamamura, T., Azuma, cinoma. Anticancer Res. 25, 709–713.
Multhoff, G. (2003a). Interaction of Hofbauer, L. C., and Schoppet, M. J., Nishimoto, N., Yoshizaki, K., Kamal, A., Thao, L., Sensintaffar, J.,
heat shock protein 70 peptide with (2004). Clinical implications of Shimoyama, T., and Kishimoto, T. Zhang, L., Boehm, M. F., Fritz, L. C.,
NK cells involves the NK receptor the osteoprotegerin/RANKL/RANK (2004). A pilot randomized trial of and Burrows, F. J. (2003). A high-
CD94. Biol. Chem. 384, 267–279. system for bone and vascular dis- a human anti-interleukin-6 recep- affinity conformation of Hsp90 con-
Gross, C., Koelch, W., DeMaio, A., eases. JAMA 292, 490–495. tor monoclonal antibody in active fers tumour selectivity on Hsp90
Arispe, N., and Multhoff, G. (2003b). Hoffman, H. M., Throne, M. L., Amar, Crohn’s disease. Gastroenterology inhibitors. Nature 425, 407–410.
Cell surface-bound heat shock pro- N. J., Sebai, M., Kivitz, A. J., 126, 989–996. Kamata, H., Honda, S., Maeda, S.,
tein 70 (Hsp70) mediates perforin- Kavanaugh, A., Weinstein, S. P., Jaiswal, M., LaRusso, N. F., and Gores, Chang, L., Hirata, H., and Karin, M.
independent apoptosis by specific Belomestnov, P., Yancopoulos, G. D., G. J. (2001). Nitric oxide in gastroin- (2005). Reactive oxygen species pro-
binding and uptake of granzyme B. Stahl, N., and Mellis, S. J. (2008). testinal epithelial cell carcinogenesis: mote TNFalpha-induced death and
J. Biol. Chem. 278, 41173–41181. Efficacy and safety of rilonacept linking inflammation to oncogene- sustained JNK activation by inhibit-
Gross, C., Schmidt-Wolf, I. G., Nagaraj, (interleukin-1 Trap) in patients with sis. Am. J. Physiol. Gastrointest. Liver ing MAP kinase phosphatases. Cell
S., Gastpar, R., Ellwart, J., Kunz- cryopyrin-associated periodic syn- Physiol. 281, G626–G634. 120, 649–661.
Schughart, L. A., and Multhoff, dromes: results from two sequential Jefferies, C. A., and O’Neill, L. A. Karin, M. (2006). Nuclear factor-
G. (2003c). Heat shock protein placebo-controlled studies. Arthritis (2000). Rac1 regulates interleukin kappaB in cancer development and
70-reactivity is associated with Rheum. 58, 2443–2452. 1-induced nuclear factor kappaB progression. Nature 441, 431–436.
increased cell surface density of Hoffmann, J., Junker, H., Schmieder, activation in an inhibitory protein Karin, M., Cao, Y., Greten, F. R., and Li,
CD94/CD56 on primary natural A., Venz, S., Brandt, R., Multhoff, kappaBalpha-independent manner Z. W. (2002). NF-kappaB in cancer:
killer cells. Cell Stress Chaperones 8, G., Falk, W., and Radons, J. by enhancing the ability of the p65 from innocent bystander to major
348–360. (2011). EGCG downregulates IL- subunit to transactivate gene expres- culprit. Nat. Rev. Cancer 2, 301–310.
Hacker, H., and Karin, M. (2006). Reg- 1RI expression and suppresses sion. J. Biol. Chem. 275, 3114–3120. Karin, M., Lawrence, T., and Nizet,
ulation and function of IKK and IL-1-induced tumorigenic factors Jendrossek, V. (2011). Targeting V. (2006). Innate immunity gone
IKK-related kinases. Sci. STKE 2006, in human pancreatic adenocarci- apoptosis pathways by cele- awry: linking microbial infections to
re13. noma cells. Biochem. Pharmacol. 82, coxib in cancer. Cancer Lett. doi: chronic inflammation and cancer.
Hanahan, D., and Weinberg, R. A. 1153–1162. 10.1016/j.canlet.2011.01.012. [Epub Cell 124, 823–835.
(2000). The hallmarks of cancer. Cell Hudson, B. I., Carter, A. M., Harja, E., ahead of print]. Kawai, T., and Akira, S. (2011). Toll-
100, 57–70. Kalea, A. Z., Arriero, M., Yang, H., Jimeno, A., Amador, M. L., Kulesza, like receptors and their crosstalk
Hanahan, D., and Weinberg, R. A. Grant, P. J., and Schmidt, A. M. P., Wang, X., Rubio-Viqueira, B., with other innate receptors in infec-
(2011). Hallmarks of cancer: the (2008). Identification, classification, Zhang, X., Chan, A., Wheelhouse, tion and immunity. Immunity 34,
next generation. Cell 144, 646–674. and expression of RAGE gene splice J., Kuramochi, H., Tanaka, K., 637–650.
Härdtner, C., Brandt, R., Multhoff, variants. FASEB J. 22, 1572–1580. Danenberg, K., Messersmith, W. A., Kawata, M., Koinuma, D., Ogami, T.,
G., and Radons, J. (2009). (-)- Hussain, S. P., and Harris, C. C. Almuete, V., Hruban, R. H., Maitra, Umezawa, K., Iwata, C., Watabe, T.,
Epigallocatechin-3-gallate compen- (2007). Inflammation and cancer: an A., Yeo, C. J., and Hidalgo, M. (2006). and Miyazono, K. (2011). TGF-beta-
sates for celecoxib-mediated up- ancient link with novel potentials. Assessment of celecoxib pharmaco- induced epithelial-mesenchymal
regulation of PGHS-2 expression Int. J. Cancer 121, 2373–2380. dynamics in pancreatic cancer. Mol. transition of A549 lung adeno-
in human pancreatic adenocarci- Hwang, W., Jung, K., Jeon, Y., Yun, S., Cancer Ther. 5, 3240–3247. carcinoma cells is enhanced by
noma cells under tumor-associated Kim, T. W., and Choi, I. (2010). Johnson, A. J., Hsu, A. L., Lin, H. P., proinflammatory cytokines derived
inflammatory conditions. Eur. J. Knockdown of the interleukin-6 Song, X., and Chen, C. S. (2002). The from RAW 264.7 macrophage cells.
Immunol. 39(Suppl. 1/09), 530–531. receptor alpha chain of dendritic cyclo-oxygenase-2 inhibitor cele- J. Biochem. doi: 10.1093/jb/mvr136.
Hartl, F. U. (1996). Molecular chap- cell vaccines enhances the therapeu- coxib perturbs intracellular calcium [Epub ahead of print].
erones in cellular protein folding. tic potential against IL-6 producing by inhibiting endoplasmic reticu- Keating, S. E., Maloney, G. M., Moran,
Nature 381, 571–579. tumors. Vaccine 29, 34–44. lum Ca2+ -ATPases: a plausible link E. M., and Bowie, A. G. (2007).

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 13


Multhoff et al. Inflammation and cancer

IRAK-2 participates in multiple toll- ONE 5, e13390. doi:10.1371/jour- H., Fletcher, F., Dunstan, C. R., of inflammatory caspases and pro-
like receptor signaling pathways to nal.pone.0013390 Lacey, D. L., and Boyle, W. J. (2000). cessing of proIL-beta. Mol. Cell 10,
NFkappaB via activation of TRAF6 Kundu, J. K., and Surh, Y. J. (2008). RANK is the intrinsic hematopoietic 417–426.
ubiquitination. J. Biol. Chem. 282, Inflammation: gearing the journey cell surface receptor that controls Matzinger, P. (1998). An innate sense
33435–33443. to cancer. Mutat. Res. 659, 15–30. osteoclastogenesis and regulation of of danger. Semin. Immunol. 10,
Khaleque, M. A., Bharti, A., Sawyer, D., Lachmann, H. J., Kone-Paut, I., bone mass and calcium metabo- 399–415.
Gong, J., Benjamin, I. J., Stevenson, Kuemmerle-Deschner, J. B., Leslie, lism. Proc. Natl. Acad. Sci. U.S.A. 97, Menu, P., and Vince, J. E. (2011).
M. A., and Calderwood, S. K. (2005). K. S., Hachulla, E., Quartier, P., 1566–1571. The NLRP3 inflammasome in health
Induction of heat shock proteins Gitton, X., Widmer, A., Patel, N., Li, X., Kim, K. W., Cho, M. L., Ju, J. and disease: the good, the bad and
by heregulin beta1 leads to protec- and Hawkins, P. N. (2009a). Use H., Kang, C. M., Oh, H. J., Min, J. the ugly. Clin. Exp. Immunol. 166,
tion from apoptosis and anchorage- of canakinumab in the cryopyrin- K., Lee, S. H., Park, S. H., and Kim, 1–15.
independent growth. Oncogene 24, associated periodic syndrome. N. H. Y. (2010). IL-23 induces recep- Merati, K., said, S. M., Andea, A., Sarkar,
6564–6573. Engl. J. Med. 360, 2416–2425. tor activator of NF-kappaB ligand F., Ben-Josef, E., Mohammad, R.,
Kim, L. S., and Kim, J. H. (2011). Lachmann, H. J., Lowe, P., Felix, S. expression in fibroblast-like syn- Philip, P., Shields, A. F., Vaitkevicius,
Heat shock protein as molecular tar- D., Rordorf, C., Leslie, K., Madhoo, oviocytes via STAT3 and NF-kappaB V., Grignon, D. J., and Adsay, N.
gets for breast cancer therapeutics. J. S., Wittkowski, H., Bek, S., Hart- signal pathways. Immunol. Lett. 127, V. (2001). Expression of inflamma-
Breast Cancer 14, 167–174. mann, N., Bosset, S., Hawkins, P. 100–107. tory modulator COX-2 in pancreatic
Kishimoto, K., Matsumoto, K., and N., and Jung, T. (2009b). In vivo Lin, W. J., and Yeh, W. C. (2005). ductal adenocarcinoma and its rela-
Ninomiya-Tsuji, J. (2000). TAK1 regulation of interleukin 1beta in Implication of Toll-like receptor and tionship to pathologic and clinical
mitogen-activated protein kinase patients with cryopyrin-associated tumor necrosis factor alpha sig- parameters. Am. J. Clin. Oncol. 24,
kinase kinase is activated by periodic syndromes. J. Exp. Med. naling in septic shock. Shock 24, 447–452.
autophosphorylation within its 206, 1029–1036. 206–209. Meylan, E., Dooley, A. L., Feldser, D.
activation loop. J. Biol. Chem. 275, Ladetto, M., Vallet, S., Trojan, A., Lin, W. W., and Karin, M. (2007). M., Shen, L., Turk, E., Ouyang,
7359–7364. Dell’Aquila, M., Monitillo, L., A cytokine-mediated link between C., and Jacks, T. (2009). Require-
Klein, B., Lu, Z. Y., Gaillard, J. P., Rosato, R., Santo, L., Drandi, D., innate immunity, inflammation, ment for NF-kappaB signalling in
Harousseau, J. L., and Bataille, R. Bertola, A., Falco, P., Cavallo, F., and cancer. J. Clin. Invest. 117, a mouse model of lung adenocarci-
(1992). Inhibiting IL-6 in human Ricca, I., De, M. F., Mantoan, B., 1175–1183. noma. Nature 462, 104–107.
multiple myeloma. Curr. Top. Micro- Bode-Lesniewska, B., Pagliano, G., Lindquist, S., and Craig, E. A. (1988). Mizel, S. B. (1982). Interleukin 1 and
biol. Immunol. 182, 237–244. Francese, R., Rocci, A., Astolfi, M., The heat-shock proteins. Annu. Rev. T cell activation. Immunol. Rev. 63,
Kobayashi, S., Werneburg, N. W., Bronk, Compagno, M., Mariani, S., Godio, Genet. 22, 631–677. 51–72.
S. F., Kaufmann, S. H., and Gores, L., Marino, L., Ruggeri, M., Omede, Lu, H., Ouyang, W., and Huang, C. Mocellin, S., Rossi, C. R., Pilati, P.,
G. J. (2005). Interleukin-6 con- P., Palumbo, A., and Boccadoro, M. (2006). Inflammation, a key event and Nitti, D. (2005). Tumor necrosis
tributes to Mcl-1 up-regulation (2005). Cyclooxygenase-2 (COX-2) in cancer development. Mol. Cancer factor, cancer and anticancer ther-
and TRAIL resistance via an Akt- is frequently expressed in multiple Res. 4, 221–233. apy. Cytokine Growth Factor Rev. 16,
signaling pathway in cholangiocar- myeloma and is an independent Lust, J. A., Lacy, M. Q., Zeldenrust, S. R., 35–53.
cinoma cells. Gastroenterology 128, predictor of poor outcome. Blood Dispenzieri, A., Gertz, M. A., Witzig, Molina, M. A., Sitja-Arnau, M.,
2054–2065. 105, 4784–4791. T. E., Kumar, S., Hayman, S. R., Rus- Lemoine, M. G., Frazier, M. L., and
Kong, G., Kim, E. K., Kim, W. S., Lee, Lambert, J. D., Hong, J.,Yang, G. Y., Liao, sell, S. J., Buadi, F. K., Geyer, S. M., Sinicrope, F. A. (1999). Increased
K. T., Lee, Y. W., Lee, J. K., Paik, J., and Yang, C. S. (2005). Inhibition Campbell, M. E., Kyle, R. A., Rajku- cyclooxygenase-2 expression in
S. W., and Rhee, J. C. (2002). Role of carcinogenesis by polyphenols: mar, S. V., Greipp, P. R., Kline, M. P., human pancreatic carcinomas and
of cyclooxygenase-2 and inducible evidence from laboratory investiga- Xiong, Y., Moon-Tasson, L. L., and cell lines: growth inhibition by
nitric oxide synthase in pancreatic tions. Am. J. Clin. Nutr. 81, 284S– Donovan, K. A. (2009). Induction nonsteroidal anti-inflammatory
cancer. J. Gastroenterol. Hepatol. 17, 291S. of a chronic disease state in patients drugs. Cancer Res. 59, 4356–4362.
914–921. Lee, H., Herrmann, A., Deng, J. H., with smoldering or indolent multi- Mori, K., Berreur, M., Blanchard,
Kortylewski, M., Kujawski, M., Wang, Kujawski, M., Niu, G., Li, Z., Forman, ple myeloma by targeting interleukin F., Chevalier, C., Guisle-Marsollier,
T., Wei, S., Zhang, S., Pilon-Thomas, S., Jove, R., Pardoll, D. M., and Yu, H. 1{beta}-induced interleukin 6 pro- I., Masson, M., Redini, F., and
S., Niu, G., Kay, H., Mule, J., Kerr, (2009). Persistently activated Stat3 duction and the myeloma prolifera- Heymann, D. (2007a). Receptor
W. G., Jove, R., Pardoll, D., and Yu, maintains constitutive NF-kappaB tive component. Mayo Clin. Proc. 84, activator of nuclear factor-kappaB
H. (2005). Inhibiting Stat3 signaling activity in tumors. Cancer Cell 15, 114–122. ligand (RANKL) directly modu-
in the hematopoietic system elicits 283–293. Maier, T. J., Schilling, K., Schmidt, lates the gene expression profile
multicomponent antitumor immu- Lee, Y. H., Ji, J. D., and Song, G. G. R., Geisslinger, G., and Grosch, S. of RANK-positive Saos-2 human
nity. Nat. Med. 11, 1314–1321. (2007). Adjusted indirect compari- (2004). Cyclooxygenase-2 (COX-2)- osteosarcoma cells. Oncol. Rep. 18,
Kulp, S. K., Yang, Y. T., Hung, son of celecoxib versus rofecoxib on dependent and -independent anti- 1365–1371.
C. C., Chen, K. F., Lai, J. P., cardiovascular risk. Rheumatol. Int. carcinogenic effects of celecoxib Mori, K., Le Goff, B., Berreur, M., Riet,
Tseng, P. H., Fowble, J. W., Ward, 27, 477–482. in human colon carcinoma cells. A., Moreau, A., Blanchard, F., Cheva-
P. J., and Chen, C. S. (2004). Lewis, A. M., Varghese, S., Xu, H., and Biochem. Pharmacol. 67, 1469–1478. lier, C., Guisle-Marsollier, I., Leger,
3-phosphoinositide-dependent pro- Alexander, H. R. (2006). Interleukin- Mantovani, A., Allavena, P., Sica, A., and J., Guicheux, J., Masson, M., Gouin,
tein kinase-1/Akt signaling repre- 1 and cancer progression: the emerg- Balkwill, F. (2008). Cancer-related F., Redini, F., and Heymann, D.
sents a major cyclooxygenase-2- ing role of interleukin-1 receptor inflammation. Nature 454, 436–444. (2007b). Human osteosarcoma cells
independent target for celecoxib in antagonist as a novel therapeutic Martin, M. U., and Wesche, H. express functional receptor activator
prostate cancer cells. Cancer Res. 64, agent in cancer treatment. J. Transl. (2002). Summary and comparison of nuclear factor-kappa B. J. Pathol.
1444–1451. Med. 4, 48. of the signaling mechanisms of 211, 555–562.
Kumar, A., Xu, J., Brady, S., Gao, H., Yu, Li, J., Sarosi, I., Yan, X. Q., Morony, S., the Toll/interleukin-1 receptor fam- Mori, T., Miyamoto, T., Yoshida,
D., Reuben, J., and Mehta, K. (2010). Capparelli, C., Tan, H. L., McCabe, ily. Biochim. Biophys. Acta 1592, H., Asakawa, M., Kawasumi, M.,
Tissue transglutaminase promotes S., Elliott, R., Scully, S., Van, G., 265–280. Kobayashi, T., Morioka, H., Chiba,
drug resistance and invasion by Kaufman, S., Juan, S. C., Sun, Y., Martinon, F., Burns, K., and Tschopp, J. K., Toyama, Y., and Yoshimura, A.
inducing mesenchymal transition Tarpley, J., Martin, L., Christensen, (2002). The inflammasome: a mole- (2011). IL-1β and TNFα-initiated
in mammary epithelial cells. PLoS K., McCabe, J., Kostenuik, P., Hsu, cular platform triggering activation IL-6-STAT3 pathway is critical in

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 14


Multhoff et al. Inflammation and cancer

mediating inflammatory cytokines DC expression of CCR7 but inhibits peritoneal mesothelial and ovarian transcription factors in regulation of
and RANKL expression in inflam- the ability of DCs to produce CCL19 cancer cells. Cytokine 7, 542–547. the heat shock response and beyond.
matory arthritis. Int. Immunol. 23, and attract naive T cells. Blood 116, Okamoto, M., Liu, W., Luo, Y., Tanaka, FASEB J. 15, 1118–1131.
701–712. 1454–1459. A., Cai, X., Norris, D. A., Dinarello, Rabe, B., Chalaris, A., May, U., Waet-
Moser, C., Schmidbauer, C., Gurtler, U., Muzio, M., Ni, J., Feng, P., and Dixit, C. A., and Fujita, M. (2010). Con- zig, G. H., Seegert, D., Williams, A.
Gross, C., Gehrmann, M., Thonigs, V. M. (1997). IRAK (Pelle) fam- stitutively active inflammasome in S., Jones, S. A., Rose-John, S., and
G., Pfister, K., and Multhoff, G. ily member IRAK-2 and MyD88 as human melanoma cells mediat- Scheller, J. (2008). Transgenic block-
(2002). Inhibition of tumor growth proximal mediators of IL-1 signal- ing autoinflammation via caspase- ade of interleukin 6 transsignaling
in mice with severe combined ing. Science 278, 1612–1615. 1 processing and secretion of abrogates inflammation. Blood 111,
immunodeficiency is mediated by Nakata, Y., Matsuda, K., Uzawa, A., interleukin-1beta. J. Biol. Chem. 285, 1021–1028.
heat shock protein 70 (Hsp70)- Nomura, M., Akashi, M., and Suzuki, 6477–6488. Radons, J., Dove, S., Neumann, D., Alt-
peptide-activated, CD94 positive G. (1995). Administration of recom- O’Neill, L. A. (2000). The interleukin- mann, R., Botzki, A., Martin, M.
natural killer cells. Cell Stress Chap- binant human IL-1 by Staphylococ- 1 receptor/Toll-like receptor super- U., and Falk, W. (2003). The inter-
erones 7, 365–373. cus enterotoxin B prevents tolerance family: signal transduction during leukin 1 (IL-1) receptor accessory
Multhoff, G., Botzler, C., Jennen, L., induction in vivo. J. Immunol. 155, inflammation and host defense. Sci. protein Toll/IL-1 receptor domain:
Schmidt, J., Ellwart, J., and Issels, R. 4231–4235. STKE 2000, re1. analysis of putative interaction sites
(1997). Heat shock protein 72 on Neckers, L., and Lee, Y. S. (2003). Can- O’Sullivan, B. J., Thomas, H. E., Pai, in vitro mutagenesis and molecu-
tumor cells: a recognition structure cer: the rules of attraction. Nature S., Santamaria, P., Iwakura, Y., Step- lar modeling. J. Biol. Chem. 278,
for natural killer cells. J. Immunol. 425, 357–359. toe, R. J., Kay, T. W., and Thomas, 49145–49153.
158, 4341–4350. Ng, P. K., Tsui, S. K., Lau, C. P., R. (2006). IL-1 beta breaks toler- Radons, J., Gabler, S., Wesche, H.,
Multhoff, G., Botzler, C., Wiesnet, M., Wong, C. H., Wong, W. H., Huang, ance through expansion of CD25+ Korherr, C., Hofmeister, R., and
Eissner, G., and Issels, R. (1995a). L., and Kumta, S. M. (2010). effector T cells. J. Immunol. 176, Falk, W. (2002). Identification of
CD3- large granular lymphocytes CCAAT/enhancer binding protein 7278–7287. essential regions in the cytoplasmic
recognize a heat-inducible immuno- beta is up-regulated in giant cell Paduch, R., Jakubowicz-Gil, J., and tail of interleukin-1 receptor acces-
genic determinant associated with tumor of bone and regulates RANKL Kandefer-Szerszen, M. (2009). sory protein critical for interleukin-
the 72-kD heat shock protein on expression. J. Cell. Biochem. 110, Expression of HSP27, HSP72 and 1 signaling. J. Biol. Chem. 277,
human sarcoma cells. Blood 86, 438–446. MRP proteins in in vitro co-culture 16456–16463.
1374–1382. Niederberger, E., Tegeder, I., Vetter, G., of colon tumour cell spheroids with Raut, C. P., Nawrocki, S., Lashinger,
Multhoff, G., Botzler, C., Wiesnet, M., Schmidtko, A., Schmidt, H., Euchen- normal cells after incubation with L. M., Davis, D. W., Khanbolooki,
Muller, E., Meier, T., Wilmanns, W., hofer, C., Brautigam, L., Grosch, S., rhTGF- beta1 and/or CPT-11. J. S., Xiong, H., Ellis, L. M., and
and Issels, R. D. (1995b). A stress- and Geisslinger, G. (2001). Cele- Biosci. 34, 927–940. McConkey, D. J. (2004). Cele-
inducible 72-kDa heat-shock pro- coxib loses its anti-inflammatory Park, H. R., Min, S. K., Cho, H. D., coxib inhibits angiogenesis by
tein (HSP72) is expressed on the sur- efficacy at high doses through acti- Kim, D. H., Shin, H. S., and Park, inducing endothelial cell apop-
face of human tumor cells, but not vation of NF-kappaB. FASEB J. 15, Y. E. (2003). Expression of osteopro- tosis in human pancreatic tumor
on normal cells. Int. J. Cancer 61, 1622–1624. tegerin and RANK ligand in breast xenografts. Cancer Biol. Ther. 3,
272–279. Ninomiya-Tsuji, J., Kishimoto, K., cancer bone metastasis. J. Korean 1217–1224.
Multhoff, G., Mizzen, L., Winchester, C. Hiyama, A., Inoue, J., Cao, Z., and Med. Sci. 18, 541–546. Ravi, R., Mookerjee, B., Bhujwalla, Z.
C., Milner, C. M., Wenk, S., Eissner, Matsumoto, K. (1999). The kinase Peters, M., Muller,A. M., and Rose-John, M., Sutter, C. H., Artemov, D.,
G., Kampinga, H. H., Laumbacher, TAK1 can activate the NIK-I kappaB S. (1998). Interleukin-6 and soluble Zeng, Q., Dillehay, L. E., Madan,
B., and Johnson, J. (1999). Heat as well as the MAP kinase cascade in interleukin-6 receptor: direct stimu- A., Semenza, G. L., and Bedi, A.
shock protein 70 (Hsp70) stimulates the IL-1 signalling pathway. Nature lation of gp130 and hematopoiesis. (2000). Regulation of tumor angio-
proliferation and cytolytic activity of 398, 252–256. Blood 92, 3495–3504. genesis by p53-induced degradation
natural killer cells. Exp. Hematol. 27, Nishimoto, N., Kanakura, Y., Aozasa, K., Pfister, K., Radons, J., Busch, R., Tid- of hypoxia-inducible factor 1alpha.
1627–1636. Johkoh, T., Nakamura, M., Nakano, ball, J. G., Pfeifer, M., Freitag, L., Genes Dev. 14, 34–44.
Multhoff, G., Pfister, K., Botzler, C., Jor- S., Nakano, N., Ikeda, Y., Sasaki, T., Feldmann, H. J., Milani, V., Issels, Reddy, S. A., Huang, J. H., and Liao,
dan, A., Scholz, R., Schmetzer, H., Nishioka, K., Hara, M., Taguchi, H., R., and Multhoff, G. (2007). Patient W. S. (1997). Phosphatidylinositol
Burgstahler, R., and Hiddemann, W. Kimura, Y., Kato, Y., Asaoku, H., survival by Hsp70 membrane phe- 3-kinase in interleukin 1 signaling.
(2000). Adoptive transfer of human Kumagai, S., Kodama, F., Nakahara, notype: association with different Physical interaction with the inter-
natural killer cells in mice with H., Hagihara, K., Yoshizaki, K., and routes of metastasis. Cancer 110, leukin 1 receptor and requirement
severe combined immunodeficiency Kishimoto, T. (2005). Humanized 926–935. in NFkappaB and AP-1 activation. J.
inhibits growth of Hsp70-expressing anti-interleukin-6 receptor antibody Pidgeon, G. P., Harmey, J. H., Kay, Biol. Chem. 272, 29167–29173.
tumors. Int. J. Cancer 88, 791–797. treatment of multicentric Castleman E., Da Costa, M., Redmond, Rose-John, S., Scheller, J., Elson,
Multhoff, G., Pfister, K., Gehrmann, M., disease. Blood 106, 2627–2632. H. P., and Bouchier-Hayes, D. G., and Jones, S. A. (2006).
Hantschel, M., Gross, C., Hafner, M., Nishimoto, N., Yoshizaki, K., Miyasaka, J. (1999). The role of endo- Interleukin-6 biology is coordinated
and Hiddemann, W. (2001). A 14- N., Yamamoto, K., Kawai, S., toxin/lipopolysaccharide in by membrane-bound and soluble
mer Hsp70 peptide stimulates nat- Takeuchi, T., Hashimoto, J., surgically induced tumour growth receptors: role in inflammation and
ural killer (NK) cell activity. Cell Azuma, J., and Kishimoto, T. in a murine model of metastatic cancer. J. Leukoc. Biol. 80, 227–236.
Stress Chaperones 6, 337–344. (2004). Treatment of rheumatoid disease. Br. J. Cancer 81, 1311–1317. Rose-John, S., and Schooltink, H.
Murshid, A., Gong, J., Stevenson, M. A., arthritis with humanized anti- Pikarsky, E., Porat, R. M., Stein, (2007). Cytokines are a thera-
and Calderwood, S. K. (2011). Heat interleukin-6 receptor antibody: a I., Abramovitch, R., Amit, S., peutic target for the prevention
shock proteins and cancer vaccines: multicenter, double-blind, placebo- Kasem, S., Gutkovich-Pyest, E., of inflammation-induced cancers.
developments in the past decade and controlled trial. Arthritis Rheum. 50, Urieli-Shoval, S., Galun, E., and Recent Results Cancer Res. 174,
chaperoning in the decade to come. 1761–1769. Ben-Neriah, Y. (2004). NF-kappaB 57–66.
Expert Rev. Vaccines 10, 1553–1568. Offner, F. A., Obrist, P., Stadlmann, functions as a tumour promoter Rudner, J., Elsaesser, S. J., Muller, A. C.,
Muthuswamy, R., Mueller-Berghaus, J., S., Feichtinger, H., Klingler, P., in inflammation-associated cancer. Belka, C., and Jendrossek, V. (2010).
Haberkorn, U., Reinhart, T. A., Herold, M., Zwierzina, H., Hittmair, Nature 431, 461–466. Differential effects of anti-apoptotic
Schadendorf, D., and Kalinski, P. A., Mikuz, G., and Abendstein, B. Pirkkala, L., Nykanen, P., and Sistonen, Bcl-2 family members Mcl-1, Bcl-
(2010). PGE(2) transiently enhances (1995). IL-6 secretion by human L. (2001). Roles of the heat shock 2, and Bcl-xL on celecoxib-induced

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 15


Multhoff et al. Inflammation and cancer

apoptosis. Biochem. Pharmacol. 79, Zhao, R., Puravs, E., Tra, J., Michael, and cardiovascular risk-where are Tang, Z. N., Zhang, F., Tang, P., Qi, X.
10–20. C. W., Misek, D. E., and Hanash, we now? Nat. Clin. Pract. Cardiovasc. W., and Jiang, J. (2011). Hypoxia
Sakoguchi-Okada, N., Takahashi- S. M. (2003). Global profiling of Med. 2, 290–300. induces RANK and RANKL expres-
Yanaga, F., Fukada, K., Shiraishi, F., the cell surface proteome of can- Stangl, S., Gehrmann, M., Riegger, sion by activating HIF-1alpha in
Taba, Y., Miwa, Y., Morimoto, S., cer cells uncovers an abundance of J., Kuhs, K., Riederer, I., Sievert, breast cancer cells. Biochem. Biophys.
Iida, M., and Sasaguri, T. (2007). proteins with chaperone function. J. W., Hube, K., Mocikat, R., Dres- Res. Commun. 408, 411–416.
Celecoxib inhibits the expression Biol. Chem. 278, 7607–7616. sel, R., Kremmer, E., Pockley, A.G., Teitelbaum, S. L. (2000). Bone resorp-
of survivin via the suppression of Shinriki, S., Jono, H., Ota, K., Ueda, M., Friedrich, L., Vigh, L., Skerra, A., tion by osteoclasts. Science 289,
promoter activity in human colon Kudo, M., Ota, T., Oike, Y., Endo, M., and Multhoff, G. (2011). Targeting 1504–1508.
cancer cells. Biochem. Pharmacol. Ibusuki, M., Hiraki, A., Nakayama, membrane heat-shock protein 70 Thiery, J. P., Acloque, H., Huang,
73, 1318–1329. H., Yoshitake, Y., Shinohara, M., (Hsp70) on tumors by cmHsp70.1 R. Y., and Nieto, M. A. (2009).
Sandau, K. B., Faus, H. G., and Brune, and Ando, Y. (2009). Humanized antibody. Proc. Natl. Acad. Sci. U.S.A. Epithelial-mesenchymal transitions
B. (2000). Induction of hypoxia- anti-interleukin-6 receptor antibody 108, 733–738. in development and disease. Cell
inducible-factor 1 by nitric oxide suppresses tumor angiogenesis and Stangl, S., Wortmann, A., Guertler, U., 139, 871–890.
is mediated via the PI 3K path- in vivo growth of human oral squa- and Multhoff, G. (2006). Control Tjiu, J. W., Liao, Y. H., Lin, S. J.,
way. Biochem. Biophys. Res. Com- mous cell carcinoma. Clin. Cancer of metastasized pancreatic carcino- Huang, Y. L., Tsai, W. L., Chu, C.
mun. 278, 263–267. Res. 15, 5426–5434. mas in SCID/beige mice with human Y., Kuo, M. L., and Jee, S. H. (2006).
Sansone, P., Piazzi, G., Paterini, P., Stril- Shirode, A. B., and Sylvester, P. W. IL-2/TKD-activated NK cells. J. Cyclooxygenase-2 overexpression in
lacci, A., Ceccarelli, C., Minni, F., (2010). Synergistic anticancer effects Immunol. 176, 6270–6276. human basal cell carcinoma cell line
Biasco, G., Chieco, P., and Bonafe, M. of combined gamma-tocotrienol Stock, A., Booth, S., and Cerundolo, V. increases antiapoptosis, angiogene-
(2009). Cyclooxygenase-2/carbonic and celecoxib treatment are associ- (2011). Prostaglandin E2 suppresses sis, and tumorigenesis. J. Invest. Der-
anhydrase-IX up-regulation pro- ated with suppression in Akt and the differentiation of retinoic acid- matol. 126, 1143–1151.
motes invasive potential and hypoxia NFkappaB signaling. Biomed. Phar- producing dendritic cells in mice Trelle, S., Reichenbach, S., Wandel, S.,
survival in colorectal cancer cells. J. macother. 64, 327–332. and humans. J. Exp. Med. 208, Hildebrand, P., Tschannen, B., Vil-
Cell. Mol. Med. 13, 3876–3887. Siddiqui, I. A., Malik, A., Adhami, V. M., 761–773. liger, P. M., Egger, M., and Juni,
Sawa, T., and Ohshima, H. (2006). Asim, M., Hafeez, B. B., Sarfaraz, S., Suganuma, M., Saha, A., and Fujiki, P. (2011). Cardiovascular safety
Nitrative DNA damage in inflamma- and Mukhtar, H. (2008). Green tea H. (2011). New cancer treatment of non-steroidal anti-inflammatory
tion and its possible role in carcino- polyphenol EGCG sensitizes human strategy using combination of green drugs: network meta-analysis. BMJ
genesis. Nitric Oxide 14, 91–100. prostate carcinoma LNCaP cells to tea catechins and anticancer drugs. 342, c7086.
Schulze, J., Albers, J., Baranowsky, TRAIL-mediated apoptosis and syn- Cancer Sci. 102, 317–323. Triantafilou, M., and Triantafilou,
A., Keller, J., Spiro, A., Stre- ergistically inhibits biomarkers asso- Sun, S. G., Lau, Y. S., Itonaga, I., Sabok- K. (2004). Heat-shock protein 70
ichert, T., Zustin, J., Amling, M., ciated with angiogenesis and metas- bar, A., and Athanasou, N. A. (2006). and heat-shock protein 90 asso-
and Schinke, T. (2010). Osteolytic tasis. Oncogene 27, 2055–2063. Bone stromal cells in pagetic bone ciate with Toll-like receptor 4 in
prostate cancer cells induce the Singh, R., Wang, B., Shirvaikar, A., and Paget’s sarcoma express RANKL response to bacterial lipopolysac-
expression of specific cytokines in Khan, S., Kamat, S., Schelling, J. R., and support human osteoclast for- charide. Biochem. Soc. Trans. 32,
bone-forming osteoblasts through Konieczkowski, M., and Sedor, J. R. mation. J. Pathol. 209, 114–120. 636–639.
a Stat3/5-dependent mechanism. (1999). The IL-1 receptor and Rho Surh, Y. J., Chun, K. S., Cha, H. H., Han, Tseng, W. W., Deganutti, A., Chen, M.
Bone 46, 524–533. directly associate to drive cell activa- S. S., Keum, Y. S., Park, K. K., and N., Saxton, R. E., and Liu, C. D.
Schwarz, R. E., Vujanovic, N. L., tion in inflammation. J. Clin. Invest. Lee, S. S. (2001). Molecular mecha- (2002). Selective cyclooxygenase-2
and Hiserodt, J. C. (1989). 103, 1561–1570. nisms underlying chemopreventive inhibitor rofecoxib (Vioxx) induces
Enhanced antimetastatic activ- Smolen, J. S., Avila, J. C., and Aletaha, activities of anti-inflammatory expression of cell cycle arrest genes
ity of lymphokine-activated killer D. (2011). Tocilizumab inhibits phytochemicals: down-regulation and slows tumor growth in human
cells purified and expanded by their progression of joint damage in of COX-2 and iNOS through pancreatic cancer. J. Gastrointest.
adherence to plastic. Cancer Res. 49, rheumatoid arthritis irrespec- suppression of NF-kappa B acti- Surg. 6, 838–843.
1441–1446. tive of its anti-inflammatory vation. Mutat. Res. 480–481, Vallabhapurapu, S., and Karin, M.
Shamovsky, I., and Nudler, E. (2008). effects: disassociation of the 243–268. (2009). Regulation and function of
New insights into the mechanism of link between inflammation Sutton, C., Brereton, C., Keogh, B., Mills, NF-kappaB transcription factors in
heat shock response activation. Cell. and destruction. Ann. Rheum. K. H., and Lavelle, E. C. (2006). A the immune system. Annu. Rev.
Mol. Life Sci. 65, 855–861. Dis. doi: 10.1136/annrheumdis- crucial role for interleukin (IL)-1 Immunol. 27, 693–733.
Shchors, K., Shchors, E., Rostker, F., 2011-200395. [Epub ahead of in the induction of IL-17-producing Virchow, R. (1863). Die krankhaften
Lawlor, E. R., Brown-Swigart, L., print]. T cells that mediate autoimmune Geschwülste. Berlin: August
and Evan, G. I. (2006). The Myc- Solomon, S. D., Wittes, J., Finn, P. V., encephalomyelitis. J. Exp. Med. 203, Hirschwald.
dependent angiogenic switch in Fowler, R., Viner, J., Bertagnolli, M. 1685–1691. Voronov, E., Carmi, Y., and Apte, R.
tumors is mediated by interleukin M., Arber, N., Levin, B., Meinert, C. Takahashi, T., Uehara, H., Bando, Y., and N. (2007). Role of IL-1-mediated
1beta. Genes Dev. 20, 2527–2538. L., Martin, B., Pater, J. L., Goss, P. E., Izumi, K. (2008). Soluble EP2 neu- inflammation in tumor angiogen-
Shimizu, M., Deguchi, A., Lim, J. Lance, P., Obara, S., Chew, E. Y., Kim, tralizes prostaglandin E2-induced esis. Adv. Exp. Med. Biol. 601,
T., Moriwaki, H., Kopelovich, J., Arndt, G., and Hawk, E. (2008). cell signaling and inhibits osteolytic 265–270.
L., and Weinstein, I. B. (2005). Cardiovascular risk of celecoxib in 6 tumor growth. Mol. Cancer Ther. 7, Vujanovic, N. L., Yasumura, S.,
(-)-Epigallocatechin gallate and randomized placebo-controlled tri- 2807–2816. Hirabayashi, H., Lin, W. C., Watkins,
polyphenon E inhibit growth and als: the cross trial safety analysis. Taketomi, A., Takenaka, K., Mat- S., Herberman, R. B., and White-
activation of the epidermal growth Circulation 117, 2104–2113. sumata, T., Shimada, M., Higashi, H., side, T. L. (1995). Antitumor
factor receptor and human epi- Sparmann, A., and Bar-Sagi, D. (2004). Shirabe, K., Itasaka, H., Adachi, E., activities of subsets of human
dermal growth factor receptor-2 Ras-induced interleukin-8 expres- Maeda, T., and Sugimachi, K. (1997). IL-2-activated natural killer cells
signaling pathways in human colon sion plays a critical role in tumor Circulating intercellular adhesion in solid tissues. J. Immunol. 154,
cancer cells. Clin. Cancer Res. 11, growth and angiogenesis. Cancer molecule-1 in patients with hepato- 281–289.
2735–2746. Cell 6, 447–458. cellular carcinoma before and after Wang, C., Deng, L., Hong, M.,
Shin, B. K., Wang, H., Yim, A. M., Le Spektor, G., and Fuster, V. (2005). Drug hepatic resection. Hepatogastroen- Akkaraju, G. R., Inoue, J., and
Naour, F., Brichory, F., Jang, J. H., insight: cyclo-oxygenase 2 inhibitors terology 44, 477–483. Chen, Z. J. (2001). TAK1 is a

Frontiers in Immunology | Inflammation January 2012 | Volume 2 | Article 98 | 16


Multhoff et al. Inflammation and cancer

ubiquitin-dependent kinase of MKK correlate in vivo. Anticancer Res. 31, Xu, X. F., Xie, C. G., Wang, X. P., E synthase is overexpressed in non-
and IKK. Nature 412, 346–351. 3273–3278. Liu, J., Yu, Y. C., Hu, H. L., and small cell lung cancer. Clin. Cancer
Wang, L., Yi, T., Kortylewski, M., Par- Wesche, H., Henzel, W. J., Shillinglaw, Guo, C. Y. (2008). Selective inhibi- Res. 7, 2669–2674.
doll, D. M., Zeng, D., and Yu, H. W., Li, S., and Cao, Z. (1997a). tion of cyclooxygenase-2 suppresses Yoshimatsu, K., Golijanin, D., Paty,
(2009a). IL-17 can promote tumor MyD88: an adapter that recruits the growth of pancreatic cancer cells P. B., Soslow, R. A., Jakobsson, P.
growth through an IL-6-Stat3 sig- IRAK to the IL-1 receptor complex. in vitro and in vivo. Tohoku J. Exp. J., DeLellis, R. A., Subbaramaiah,
naling pathway. J. Exp. Med. 206, Immunity 7, 837–847. Med. 215, 149–157. K., and Dannenberg, A. J. (2001b).
1457–1464. Wesche, H., Korherr, C., Kracht, M., Yamada, T., Tsuda, M., Takahashi, T., Inducible microsomal prostaglandin
Wang, S., Liu, Q., Zhang, Y., Liu, K., Yu, Falk,W., Resch, K., and Martin, M. U. Totsuka, Y., Shindoh, M., and Ohba, E synthase is overexpressed in col-
P., Liu, K., Luan, J., Duan, H., Lu, Z., (1997b). The interleukin-1 receptor Y. (2011). RANKL expression specif- orectal adenomas and cancer. Clin.
Wang, F., Wu, E., Yagasaki, K., and accessory protein (IL-1RAcP) is ically observed in vivo promotes Cancer Res. 7, 3971–3976.
Zhang, G. (2009b). Suppression of essential for IL-1-induced activation epithelial mesenchymal transition Yu, H., Kortylewski, M., and Pardoll,
growth, migration and invasion of of interleukin-1 receptor-associated and tumor progression. Am. J. D. (2007). Crosstalk between cancer
highly-metastatic human breast can- kinase (IRAK) and stress-activated Pathol. 178, 2845–2856. and immune cells: role of STAT3 in
cer cells by berbamine and its mol- protein kinases (SAP kinases). J. Biol. Yamashita, M., Kurokawa, K., Sato, the tumour microenvironment. Nat.
ecular mechanisms of action. Mol. Chem. 272, 7727–7731. Y., Yamagata, A., Mimura, H., Rev. Immunol. 7, 41–51.
Cancer 8, 81. Wheeler, D. S., Chase, M. A., Senft, A. Yoshikawa, A., Sato, K., Nakano, A., Yu, H., Pardoll, D., and Jove, R. (2009).
Watters, T. M., Kenny, E. F., and O’Neill, P., Poynter, S. E., Wong, H. R., and and Fukai, S. (2010). Structural basis STATs in cancer inflammation and
L. A. (2007). Structure, function and Page, K. (2009). Extracellular Hsp72, for the Rho- and phosphoinositide- immunity: a leading role for STAT3.
regulation of the Toll/IL-1 receptor an endogenous DAMP, is released by dependent localization of the exo- Nat. Rev. Cancer 9, 798–809.
adaptor proteins. Immunol. Cell Biol. virally infected airway epithelial cells cyst subunit Sec3. Nat. Struct. Mol. Zeisberg, M., and Neilson, E. G.
85, 411–419. and activates neutrophils via Toll- Biol. 17, 180–186. (2009). Biomarkers for epithelial-
Weaver, C. T., Hawrylowicz, C. M., like receptor (TLR)-4. Respir. Res. Yang, J., and Weinberg, R. A. (2008). mesenchymal transitions. J. Clin.
and Unanue, E. R. (1988). T helper 10, 31. Epithelial-mesenchymal transition: Invest. 119, 1429–1437.
cell subsets require the expression Whiteside, T. L., Vujanovic, N. L., at the crossroads of development Zumsteg, A., and Christofori, G. (2009).
of distinct costimulatory signals by and Herberman, R. B. (1998). Nat- and tumor metastasis. Dev. Cell 14, Corrupt policemen: inflammatory
antigen-presenting cells. Proc. Natl. ural killer cells and tumor therapy. 818–829. cells promote tumor angiogenesis.
Acad. Sci. U.S.A. 85, 8181–8185. Curr. Top. Microbiol. Immunol. 230, Yasumura, S., Lin, W. C., Hirabayashi, Curr. Opin. Oncol. 21, 60–70.
Wei, D., Wang, L., He, Y., Xiong, H. 221–244. H., Vujanovic, N. L., Herberman,
Q., Abbruzzese, J. L., and Xie, K. Wilson, K. T., Fu, S., Ramanujam, K. S., R. B., and Whiteside, T. L. (1994). Conflict of Interest Statement: The
(2004). Celecoxib inhibits vascular and Meltzer, S. J. (1998). Increased Immunotherapy of liver metas- authors declare that the research was
endothelial growth factor expres- expression of inducible nitric oxide tases of human gastric carcinoma conducted in the absence of any com-
sion in and reduces angiogenesis and synthase and cyclooxygenase-2 in with interleukin 2-activated nat- mercial or financial relationships that
metastasis of human pancreatic can- Barrett’s esophagus and associated ural killer cells. Cancer Res. 54, could be construed as a potential con-
cer via suppression of Sp1 transcrip- adenocarcinomas. Cancer Res. 58, 3808–3816. flict of interest.
tion factor activity. Cancer Res. 64, 2929–2934. Yip-Schneider, M. T., Sweeney, C. J.,
2030–2038. Wink, D. A.,Vodovotz,Y., Laval, J., Laval, Jung, S. H., Crowell, P. L., and Mar- Received: 25 October 2011; accepted: 28
Weichert, W., Boehm, M., Gekeler, V., F., Dewhirst, M. W., and Mitchell, J. shall, M. S. (2001). Cell cycle effects December 2011; published online: 12 Jan-
Bahra, M., Langrehr, J., Neuhaus, B. (1998). The multifaceted roles of of nonsteroidal anti-inflammatory uary 2012.
P., Denkert, C., Imre, G., Weller, C., nitric oxide in cancer. Carcinogenesis drugs and enhanced growth inhi- Citation: Multhoff G, Molls M and
Hofmann, H. P., Niesporek, S., Jacob, 19, 711–721. bition in combination with gem- Radons J (2012) Chronic inflammation
J., Dietel, M., Scheidereit, C., and Wu, S., Rhee, K. J., Albesiano, E., citabine in pancreatic carcinoma in cancer development. Front. Immun.
Kristiansen, G. (2007). High expres- Rabizadeh, S., Wu, X., Yen, H. R., cells. J. Pharmacol. Exp. Ther. 298, 2:98. doi: 10.3389/fimmu.2011.00098
sion of RelA/p65 is associated with Huso, D. L., Brancati, F. L., Wick, 976–985. This article was submitted to Frontiers in
activation of nuclear factor-kappaB- E., McAllister, F., Housseau, F., Par- Yokota, S., Miyamae, T., Imagawa, T., Inflammation, a specialty of Frontiers in
dependent signaling in pancreatic doll, D. M., and Sears, C. L. (2009). Iwata, N., Katakura, S., and Mori, Immunology.
cancer and marks a patient popu- A human colonic commensal pro- M. (2004). Inflammatory cytokines Copyright © 2012 Multhoff, Molls and
lation with poor prognosis. Br. J. motes colon tumorigenesis via acti- and systemic-onset juvenile idio- Radons. This is an open-access article
Cancer 97, 523–530. vation of T helper type 17 T cell pathic arthritis. Mod. Rheumatol. 14, distributed under the terms of the Cre-
Weiss, T. W., Simak, R., Kaun, C., Rega, responses. Nat. Med. 15, 1016–1022. 12–17. ative Commons Attribution Non Com-
G., Pfluger, H., Maurer, G., Huber, Xiong, H. Q. (2004). Molecular tar- Yoshimatsu, K.,Altorki, N. K., Golijanin, mercial License, which permits non-
K., and Wojta, J. (2011). Onco- geting therapy for pancreatic can- D., Zhang, F., Jakobsson, P. J., Dan- commercial use, distribution, and repro-
statin M and IL-6 induce u-PA and cer. Cancer Chemother. Pharmacol. nenberg, A. J., and Subbaramaiah, duction in other forums, provided the
VEGF in prostate cancer cells and 54(Suppl. 1), S69–S77. K. (2001a). Inducible prostaglandin original authors and source are credited.

www.frontiersin.org January 2012 | Volume 2 | Article 98 | 17

You might also like